118
Multiscale modelling of snow depth over an agricultural field in a small catchement in southern ontario, canada. by Robert Neilly A thesis presented to the University of Waterloo in fulfillment of the thesis requirement for the degree of Master of Science in Geography Waterloo, Ontario, Canada, 2011 © Robert Neilly 2011

Multiscale modelling of snow depth over an agricultural

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Multiscale modelling of snow depth over an agricultural

Multiscale modelling of snow depth over an agricultural field in a small

catchement in southern ontario, canada.

by

Robert Neilly

A thesis

presented to the University of Waterloo

in fulfillment of the

thesis requirement for the degree of

Master of Science

in

Geography

Waterloo, Ontario, Canada, 2011

© Robert Neilly 2011

Page 2: Multiscale modelling of snow depth over an agricultural

ii

I hereby declare that I am the sole author of this thesis. This is a true copy of the thesis, including

any required final revisions, as accepted by my examiners.

I understand that my thesis may be made electronically available to the public.

Page 3: Multiscale modelling of snow depth over an agricultural

iii

Abstract

Snow is a common overlying surface during winter-time and the redistribution of snow by wind

is a very important concept for any hydrological research project located within the cryosphere.

Wind redistributes snow by eroding it from areas of high wind speed, such as ridge tops and

windward slopes, and deposits it in areas of lower wind speeds, such as the lees of ridge tops,

vegetation stands, and topographic depressions. The accurate modelling of blowing snow

processes such as erosion, deposition, and sublimation have proven to be rather problematic. The

largest issue that many modellers must deal with is the accurate collection of solid precipitation

throughout the winter season. Without this, incorrect energy and mass balances can occur. This

thesis makes use of a new method of acquiring solid precipitation values through the use of an

SR50a ultrasonic snow depth sensor and then incorporates it into a version of the Cold Regions

Hydrological Model (CRHM) which includes the Prairie Blowing Snow Model (PBSM) and the

Minimal Snowmelt Model (MSM) modules. The model is used to simulate seasonal snow depth

over an agricultural field in southern Ontario, Canada and is driven with half-hourly locally

acquired meteorological data for 83 days during the 2008-2009 winter season. Semi-automated

snow surveys are conducted throughout the winter season and the collected in situ snow depth

values are compared to the simulated snow depth values at multiple scales. Two modelling

approaches are taken to temporally and spatially test model performance. A lumped approach

tests the model‟s ability to simulate snow depth from a small point scale and from a larger field

scale. A distributed approach separates the entire field site into three hydrological response units

(HRUs) and tests the model‟s ability to spatially discretize at the field scale. HRUs are

differentiated by varying vegetation heights throughout the field site. Temporal analysis

compares the simulated results to each day of snow survey and for the entire field season. Model

performance is statistically analyzed through the use of a Root Mean Square Difference

(RMSD), Nash-Sutcliffe coefficient (NS), and Model Bias (MB). Both the lumped and

distributed modelling approaches fail to simulate the early on-set of snow but once the snow-

holding capacities are reached within the field site the model does well to simulate the average

snow depth during the latter few days of snow survey as well as throughout the entire field

season. Several model limitations are present which prevent the model from incorporating the

scaling effects of topography, vegetation, and man-made objects as well as the effects from

certain energy fluxes. These limitations are discussed further.

Page 4: Multiscale modelling of snow depth over an agricultural

iv

Acknowledgements

There are many people that I would like to thank that made it possible to complete this thesis.

First and foremost I would like to thank my advisor Dr. Richard Kelly who took a chance on me

and allowed me to come back to the University of Waterloo to join his excellent team of snow

scientists. His passion for the subject inspired me to want to learn as much as I could during my

tenure under his guidance. I especially want to thank him for allowing me the freedom to create

my own study. It may have been more time consuming and at times, painful but in the end it was

a great learning experience that I will never forget. Most importantly however, I would like to

thank him for his enduring patience from start to finish of this little project. First round‟s on me!

I‟d also like to thank Dr. Claude Duguay who provided me with the opportunity to gain some

valuable and memorable field experiences outside of my own study. I‟ll never forget our little

Alaskan snowmachine excursion in -65C weather. At Wilfrid Laurier University I would like to

thank Dr. Bill Quinton who introduced me to the Cold Regions Hydrological Model and helped

peak my interest in its applicability to modelling snow depth. I‟d also like to thank Dr. Mike

English, whose long-standing local connections at Strawberry Creek helped us establish contact

with the Zinger family who granted us permission to set up on Martz farm. At the University of

Saskatchewan I would like to thank Mr. Tom Brown and Mr. Matt Macdonald for their help and

support when I had a whole host of modelling questions. I feel as if it was a learning experience

for us all.

Secondly, but equally as important, I would like to thank my family for their support through this

entire process. They saw firsthand the process that I went through in order to complete this thesis

and they were there to listen to all the issues even when most of the time they didn‟t really

understand what I was talking about. They were also there for me during all „the other stuff‟ that

Page 5: Multiscale modelling of snow depth over an agricultural

v

went on in my life and for that, I am very grateful. It was nice to know that I could always count

on them for support and to put a roof over my head or pay the odd bill when I needed it.

Lastly I would like to thank my friends and fellow students who graciously accepted the change

in topic whenever they would ask “are you done yet?”

Page 6: Multiscale modelling of snow depth over an agricultural

vi

Table of Contents

AUTHOR’S DECLARATION ii

ABSTRACT iii

ACKNOWLEDGEMENTS iv

TABLE OF CONTENTS vi

LIST OF FIGURES ix

LIST OF TABLES x

LIST OF EQUATIONS xi

LIST OF SYMBOLS xiii

1. INTRODUCTION 1

1.1 THE NATURE OF SNOW IN THE WATER CYCLE 1

1.2 HYDROLOGICAL MODELLING 3

1.2.1 Small Scale Models 5

1.2.2 Regional Snow Models 5

1.2.3 Global Climate Models 6

1.2.4 Model Challenges 7

1.3 OBJECTIVES 9

1.4 THESIS LAYOUT 10

2. BACKGROUND 11

2.1 PHYSICAL PROCESSES OF SNOW 12

2.1.1 Precipitation 12

2.1.2 Density 13

2.1.3 Snow Depth 15

2.1.4 Snow Water Equivalent 15

2.2 ENERGY AND MASS BALANCE COMPONENT MODELS FOR A SNOWPACK 17

2.2.1 Energy Balance components 17

2.2.2 Shortwave Radiation Input (K) 18

2.2.3 Long-wave Radiation Input (L) 21

2.2.4 Sensible Heat Exchange with the Atmosphere (H) 22

2.2.5 Latent Heat Exchange with the Atmosphere (LE) 23

2.2.6 Heat Input from Rain (R) 23

2.2.7 Heat Input from the Ground 24

2.2.8 A comparison of Four Different Energy Balance Equations 24

2.2.9 Mass Balance Components 26

2.2.10 Sublimation, Evaporation, and Condensation of a Snowpack 27

2.2.11 Canopy Interception 27

2.2.12 Blowing Snow 28

2.3 BLOWING SNOW AND THE WINTER MOISTURE BUDGET 28

Page 7: Multiscale modelling of snow depth over an agricultural

vii

2.3.1 The winter moisture budget 30

2.3.1.i Precipitation 30

2.3.1.ii Snow depth measurements 31

2.3.1.iii Sublimation 32

2.4 Summary of Energy Balance + Mass Balance of Snow 33

3. THE MODELS 35

3.1 THE PRAIRIE BLOWING SNOW MODEL 35

3.1.1 Saltation 35

3.1.2 Suspension 37

3.1.3 Sublimation 39

3.1.4 Erosion and Deposition 41

3.2 THE MINIMAL SNOWMELT MODEL 44

3.3 SUMMARY 45

4. METHODS 47

4.1 STUDY SITE 47

4.2 DIGITAL IMAGERY AND DATA 51

4.3 METEOROLOGICAL DATA 52

4.3.1 Ultrasonic snow depth sensors vs. traditional snow measurement techniques 53

4.3.2 Meteorological data alteration in order for proper model runs 56

4.4 SNOW SURVEYS 56

4.5 MODEL RUNS 58

4.5.1 Lumped Model – Point Scale 58

4.5.2 Lumped Model – Field Scale 59

4.5.3 Distributed Model – Field Scale 59

4.5.4 Model Parameters 59

4.6 MODEL TESTING AND VERIFICATION 64

4.7 SUMMARY 67

5. RESULTS 68

5.1 FIELD RESULTS 68

5.2 MODEL RESULTS 70

5.3 COMPARISON OF SIMULATED MODEL RESULTS AND OBSERVED

SNOW SURVEY RESULTS AT THE POINT AND FIELD SCALES. 72

5.3.1 Point Scale Results: Lumped 73

5.3.2 Field Scale Results: Lumped 74

5.3.3 Field Scale Results: Distributed 75

5.4 SUMMARY 76

6. DISCUSSION AND CONCLUSIONS 78

6.1 EVALUATING MODEL PERFORMANCE AND SOURCES OF ERROR 78

6.2 FUTURE CONSIDERATIONS 83

Page 8: Multiscale modelling of snow depth over an agricultural

viii

6.3 CONCLUSIONS 84

REFERENCES 86

APPENDICES 94

APPENDIX 1 94

APPENDIX 2 96

Page 9: Multiscale modelling of snow depth over an agricultural

ix

List of Figures

1-1 Classification of hydrological modelling by means of how they deal with the

factors of randomness, space, and time 4

1-2 SnowMIP2 analysis from Boulder, CO 7

2-1 A block of snow 17

2-2 Exponential decay of solar radiation with depth for a typical snowpack or glacier 20

2-3 Schematic diagram of the mass balance of a snowpack 26

3-1 Process of saltation and suspension of sediments into the atmosphere 36

3-2 Cross sectional view of a column of blowing snow 42

4-1 Location of Strawberry Creek field site 47

4-2 360 view of field sight from eight cardinal directions 49

4-3 Grand River watershed 50

4-4 In situ view of meteorological tower 53

4-5 Seasonal snow depth measurements from SR50a compared to manual snow depth

measurements from the Waterloo-Wellington International Airport 55

4-6 Map of snow depth transects 57

4-7 Distributed map of CRHM hydrological response units 60

4-8 Simplified schematic diagram of the progression of module observations through

the CRHM model modules 63

5-1 Corrected meteorological data from Strawberry Creek 69

5-2 Comparison of averaged model and observed snow depth 71

5-3 Comparison of averaged model and observed snow depth for each HRU in the

distributed modelling approach 72

Page 10: Multiscale modelling of snow depth over an agricultural

x

List of Tables

1-1 Classification of snow 2

2-1 Classification scheme of SWE 16

4-1 List of selected meteorological instrumentation 52

4-2 CRHM model parameters 61

4-3 Characteristics of HRU parameters for blowing snow 62

5-1 Monthly climatic averages for Strawberry Creek (2008-2009 model season) and

Waterloo-Wellington International Airport (1971-2000) 68

5-2 Sample size of snow depth from each day of snow survey used for statistical

analysis 73

5-3 RMSD, NS, and MB results for the lumped modelling approach at the point scale 73

5-4 RMSD, NS, and MB results for the lumped modelling approach at the field scale 74

5-5 RMSD, NS, MB results for the distributed modelling approach at the field scale 75

Page 11: Multiscale modelling of snow depth over an agricultural

xi

List of Equations

[2.1] Fresh snow density 1 14

[2.2] Fresh snow density 2 14

[2.3] Snow water equivalent (SWE) 15

[2.4] Simple equation to explain energy balance 17

[2.5] Net rate of energy exchange 18

[2.6] Net shortwave radiation 18

[2.7] Transmission, absorption, and reflectance of shortwave radiation 19

[2.8] Beer‟s Law 20

[2.9] Net longwave radiation 21

[2.10] Net longwave radiation under full cloud cover 22

[2.11] Sensible heat exchange with the atmosphere 22

[2.12] Sensible heat equation adapted for snow 22

[2.13] Latent heat exchange with the atmosphere 23

[2.14a] Heat input from the rain for a „wet‟ snowpack 23

[2.14b] Heat input from the rain for a „cold‟ snowpack 23

[2.15] Heat input from the ground 24

[2.16] Marsh‟s energy balance equation 24

[2.17] Male‟s energy balance equation 1 25

[2.18] Male‟s energy balance equation 2 25

[2.19] Net change in mass storage 26

[2.20] Snowfall canopy interception 27

[2.21] Maximum snowload on a canopy 28

[2.22] The winter moisture budget 30

[3.1] Saltation transportation 36

[3.2] Atmospheric friction velocity 36

[3.3] Roughness height for blowing snow 37

Page 12: Multiscale modelling of snow depth over an agricultural

xii

[3.4] Suspension transportation rate 38

[3.5] Upper boundary height of a column of blowing snow 38

[3.6] Lower boundary height of a column of blowing snow 38

[3.7] Rate of change of mass with respect to time 39

[3.8] Nusselt and Sherwood Numbers 40

[3.9] Reynold‟s Number 40

[3.10] Ventilation velocity in the suspension layer 40

[3.11] Ventilation velocity in the saltation layer 40

[3.12] Undersaturation of water vapour with decreasing relative humidity with

increasing height 41

[3.13] Undersaturation of water vapour with increasing relative humidity with

increasing height 41

[3.14] Rate of sublimation 41

[3.15] Rate of erosion/deposition 42

[3.16] Mass balance equation for an HRU 43

[3.17] Sensible heat flux within the MSM module 44

[3.18] Latent heat flux within the MSM module 44

[3.19] Net radiation absorbed by snow surface within the MSM module 45

[3.20] Energy balance equation within the MSM module 45

[3.21] Mass balance equation within the MSM module 45

[4.1] Speed of sound in air 54

[4.2] Root mean square difference 65

[4.3] Nash-Sutcliffe coefficient 65

[4.4] Model bias 65

Page 13: Multiscale modelling of snow depth over an agricultural

xiii

List of Symbols

Symbol Quantity SI Units

Roman capital letters

A surface area m3

Ast exposed silhouette area of a single vegetation element

CH surface exchange coefficient

Ch boundary layer coefficient m-0.27

s1.27

Csalt saltation constant m s-1

Cst coefficient for vegetation

Cz coefficient for prairie snowcover

D snow water equivalent depth mm

diffusivity of water m2s

-1

Dp Snow distribution factor

DH diffusivity of sensible heat m2s

-1

DM diffusivity of momentum in turbulent air m2s

-1

DWV diffusivity of water vapour in turbulent air m2s

-1

E evapotranspiration mm; kg m-2

s-1

EB blowing snow sublimation rate kg m-2

s-1

F downwind transport rate kg m-2

s-1

G heat input from the ground (substrate) W m-2

H sensible heat exchange with the atmosphere W m-2

I canopy interception kg m-2

I* maximum snow load kg m-2

I0 snow load kg m-2

K shortwave radiation input W m-2

K↓ incoming shortwave radiation W m-2

K↑ outgoing shortwave radiation W m-2

Page 14: Multiscale modelling of snow depth over an agricultural

xiv

Ka Thermal conductivity of air W m-1

K-1

L* longwave radiation exchange W m

-2

L↓ incoming longwave radiation W m-2

L↑ outgoing longwave radiation W m-2

LAI leaf area index

LE latent heat exchange with the atmosphere W m-2

M molecular weight of water kg kmol-1

melt rate kg m-2

s-1

MB Model Bias

N total number of particles per unit volume

NS Nash-Sutcliffe coefficient

Nst number of vegetation elements per unit area m2

Nu Nusselt number

P air pressure; Pa

precipitation

snowfall rate kg m-2

s-1

Q* net all-wave energy W m-2

Q1 specific humidity at the measurement height

QG heat flux from the ground at the soil-snow interface W m-2

QH sensible heat flux W m-2

Qi energy flux due to net radiation, sensible and latent heat fluxes

and heat transfers from lower layers of the snowpack W m-2

QLE latent heat flux W m-2

QM available energy for melt W m-2

Qr radiant energy received by a particle W m-2

Qsat saturation humidity at the snow surface

Qsalt saltation transportation rate kg m-1

s-1

Page 15: Multiscale modelling of snow depth over an agricultural

xv

Qsubl sublimation rate per unit area of snowcover kg m-2

s-1

Qsusp suspension transportation rate kg m-1

s-1

R heat input from rain W m-2

universal gas constant J kmol-1

K-1

Re Reynolds number

RMSD Root mean square deviation (same as obs/simulated)

Rs snowfall rate for a unit of time mm s-1

S net rate of energy exchanges; W m-2

Sublimation

SD snow depth m

Sh Sherwood number

Sp tree species coefficient

SWE snow water equivalent mm

Ta temperature of air K, (°C)

Tat absolute radiative temperature of the atmosphere and canopy °C

Tc cloud base temperature °C

Tr temperature of the rain °C

Ts surface temperature °C

Tss absolute temperature of the snow surface °C

Roman small letters

a extinction coefficient m-1

ca heat capacity of air at a constant temperature J kg-1

°C-1

ccan canopy coverage fraction

csubl sublimation loss rate coefficient

csuc snow unloading coefficient

cw heat capacity of water J kg-1

°C-1

Page 16: Multiscale modelling of snow depth over an agricultural

xvi

dr particle size increment m

ds snow depth cm

dSWE/dt surface snow accumulation rate kg m-2

s-1

dT/dz temperature gradient of the soil

du/dt rate of change of internal energy KJ s-1

m-2

e base of Naperian logarithms

ei saturation vapour pressure over ice Pa

es vapour pressure at the surface Pa

g acceleration due to gravity m s-1

h specific enthalpy KJ kg-1

h* lower boundary of suspension layer

hs height of snow m

k von Karman’s constant m

kG thermal conductivity of the soil W m-1

°C-1

m measurement height m

mass kg

n sample size

n(z) mass concentration of suspended snow kg m-3

p precipitation mm

probability of blowing snow occurrence within the HRU

phi energy transfer due to precipitation kg m-2

s-1

qv vertical flux at the top of the controlled volume kg m-2

s-1

r rate of rainfall mm d-1

radius of snow particle contacting mass

u velocity of wind m s-1

u* atmospheric friction velocity m s-1

v kinematic viscosity of air m2 s

-1

Page 17: Multiscale modelling of snow depth over an agricultural

xvii

vr ventilation velocity m s-1

ws terminal velocity m s-1

z vertical distance m

zb upper boundary of suspension layer

Greek capital letters

Δt change in time s

ΔQ change in heat energy absorbed by the snowpack W m-2

Δr net surface and subsurface horizontal exchange mm

ΔS net change in mass storage per unit volume kg; kg m-3

s-1

(per unit horizontal area) kg m-3

s-1

Ψ transmissivity

Greek small letters

α surface albedo

ε surface emissivity

εss emissivity of the snow surface

εat integrated emissivity of the atmosphere and canopy

ζ absorptivity

λf latent heat of fusion J kg-1

λs latent heat of sublimation J kg-1

λv latent heat of vapourization J kg-1

σ Stefan-Boltzmann constant W m-2

K-4

water vapour deficit with respect to ice

σ2 undersaturation of water vapour

ρa density of air kg m-3

ρs density of snow kg m-3

ρsat saturation density of water vapour kg m-3

ρw density of water kg m-3

Page 18: Multiscale modelling of snow depth over an agricultural

xviii

Subscripts

a air

n sheer stress applied to non-erodible surfaces

o surface

observed values

s simulated results

t sheer stress applied to snow surfaces

Page 19: Multiscale modelling of snow depth over an agricultural

1

Chapter 1.

Introduction

1.1 The Nature of Snow in the Water Cycle

Snow exerts a major influence on the environment and human activity on the Earth‟s surface.

Through its high albedo, it plays a significant role in the climate system and is a sensitive

indicator of climate change and variability. There are many locations in the world that are

seasonally covered with snow and people living in these locations have learned to adapt to its

many adverse and beneficial effects. Blowing snow on transportation, heavy snow loads on

structures and buildings, rapid snowmelt resulting in flooding (especially over frozen ground),

and acid-shock entering aquatic systems are just some of the examples of adverse effects of snow

(Hiemstra et al., 2002). Snow removal is a prominent feature in municipal budgeting and more

often than not, large snow storms can have a significant effect on a population‟s mobility due to

the disruption of power and transportation services. These same storms can also lead to an

increased amount of shovelling-related heart attacks, resulting in numerous cases of

hospitalization and even fatalities each year. Examples of beneficial effects include the fact that

snow acts as an insulator to flora and fauna, providing shelter from the harsh winter climates.

Melting snow augments soil water reserves and replenishes ground water supplies and reservoirs.

This is very important to agricultural activities in southern Ontario as many farms rely on the

melting of the seasonal snow cover to aid in ground water recharge, decreasing the need for

increased irrigation. Furthermore, the winter recreation industry relies heavily on seasonal snow

cover. In southern Ontario alone there are numerous downhill and cross-country ski centres, as

well as thousands of kilometres of groomed snowmobile trails for recreation.

Page 20: Multiscale modelling of snow depth over an agricultural

2

Patterns of snow accumulation are not uniform across the globe. Its physical nature varies from

location to location and is highly dependent on latitude, altitude, and meteorological conditions

such as temperature, wind, relative humidity, and vapour pressure. Table 1-1 demonstrates a

simple classification of snow derived by Sturm et al. (1995) from an analysis of snow cover

meteorological conditions in Alaska.

Snow Cover Class Description

Tundra

A thin, cold, windblown snow. Maximum depth, ~75 cm. Usually found above

or north of tree line. Consists of a basal layer of depth hoar overlain by multiple

wind slabs. Surface zastrugi common. Melt features rare.

Taiga

A thin to moderately deep low-density cold snowcover. Maximum depth 120

cm. Found in cold climates in forests where wind, initial snow density, and

average winter air temperatures are all low. By late winter consists of 50-80%

depth hoar covered by low-density new snow.

Alpine

An intermediate to cold deep snowcover. Maximum depth, ~250 cm. Often

alternate thick and thin layers, common as well as occasional wind crusts. Most

new snowfalls are low density. Melt features occur but are generally

insignificant.

Maritime

A warm deep snowcover. Maximum depth can be in excess of 300 cm. Melt

features (ice layers, percolation columns) very common. Coarse-grained snow

due to wetting ubiquitous. Basal melting common.

Ephemeral

A thin, extremely warm snowcover. Ranges from 0 to 50 cm. Shortly after basal

melting common. Melt features common. Often consist of a single snowfall,

which melts away; then a new snowcover re-forms at the next snowfall.

Prairie A thin (except in drifts) moderately cold snowcover with substantial wind

drifting. Maximum depth, ~100 cm. Wind slabs and drifts common.

Mountain, special

cases

A highly variable snowcover, depending on solar radiation effects and local

wind patterns. Usually deeper than associated type of snowcover from adjacent

lowlands. Table 1-1. Classification of snow cover (Sturm et al., 1995)

Even though the different classes of snow cover were derived from data acquired in Alaska, they

are transferable to many locations. Southern Ontario experiences both ephemeral and prairie-like

snow covers depending on the time of year and climatological conditions. In terms of

hydrological modelling, the understanding of specific types of snow classes is important because

it can lead to appropriate parameterization in the modelled location.

Page 21: Multiscale modelling of snow depth over an agricultural

3

This brief overview of snow highlights some of the environmental, economical, social, and

climatological aspects that snow has on our lives. Further information of the physical properties

of snow and its importance to energy and mass balance models can be found in Chapter 2.

1.2 Hydrological Modelling

Estimating snow accumulation from point in situ measurements and hydrologic models is an

important approach to help manage snow resources. To accurately predict or estimate a

hydrological process it is important to understand the nature of a system. A system is essentially

a set of equations linking the input and the output variables. These variables can be functions of

space, time, and whether or not there is an effect of randomness. Hydrological models are

“simplified models that are used to represent real-life systems and may be substitutes of the real

systems for certain purposes” (Diskin, 1970). To develop a model with time and space varying

parameter and variables is challenging. From a practical point of view it becomes necessary to

simplify the model by neglecting some of the sources of variation. Chow et al (1988) developed

a method of classifying hydrological models according to how they deal with these factors of

variation. This classification scheme (Figure 1-1) is created by asking three simple questions:

Will the model variables be random or not? Will they vary or be uniform in space? Will they

vary or be constant in time?

Page 22: Multiscale modelling of snow depth over an agricultural

4

Figure 1-1: Classification of hydrological modelling by the means of how they deal with the factors of randomness, space,

and time. Modelled after Chow et al, 1988.

The first line in the classification scheme deals with how the model relates to the factor of

randomness. Although all hydrological phenomena often have some form of randomness, the

resulting variability in the model output may be quite small when compared to the variability

from known factors. When this is the case, a deterministic model is appropriate because this type

of model does not consider randomness. In a deterministic model, a given input always produces

the same output. Stochastic models on the other hand have outputs that are at least partially

random. When the random variation is large, a stochastic model is more appropriate due to the

fact that the actual output from such a model could be quite different from the output produced

by a deterministic model. This thesis uses a deterministic model, therefore, the following section

from this point forward deals solely with deterministic modelling rather than stochastic. For

more information concerning stochastic modelling, see Chow et al, (1988). It is important to note

that the above statements assume a normal distribution to the observed data. In some cases

however, a poor fit to observed data is obtained and a different distribution function should be

considered. As a hydrological example, if the probability of a certain number of events occurring

System

f (randomness, space, time)

Deterministic

Lumped

Steady FlowUnsteady

Flow

Distributed

Steady FlowUnsteady

Flow

Stochastic

Space-independent

Time-independent

Time-correlated

Space-correlated

Time-independent

Time-correlated

OutputInput

Page 23: Multiscale modelling of snow depth over an agricultural

5

in a specified period of time is needed (e.g. when the temperature exceed a certain amount or

certain levels of snowfall are reached), then a Poisson and/or Gamma distribution are frequently

used.

The second line in the classification scheme deals with how the model relates to spatial variation.

Real life hydrological problems deal with space in all three dimensions but when it comes to

developing a model to practically represent such conditions, it can be quite cumbersome. A

deterministic lumped model spatially averages the system so that it is represented as a single

point in space without dimensions. Contrary to this, a deterministic distributed model defines the

model variables as functions of space by considering the hydrological process to be occurring at

various points in space.

The third line in the classification scheme deals with how the model relates to the variability of

time. Deterministic models can be broken down into either steady-flow models, where the flow

rate varies with time, and unsteady-flow.

Many hydrological models can provide the user with an accurate simulation over the required

domain. These models are generally quite simple and may in fact only be able to simulate a few

processes and tend to be rather small in scale. When attempting to increase the scale and/or the

number of processes being modelled, many hydrological models are coupled with land surface

models as well as atmospheric circulation models. These models range from being simple to

complex.

1.2.1 Small Scale Models

Most of the smaller scale snow models are one-dimensional energy balance models that deal

with a snow pack over a very small area. Some of the more complicated models venture into two

Page 24: Multiscale modelling of snow depth over an agricultural

6

or three dimensions but are more susceptible to error due to an increase in the amount of

variables involved. Some examples of such models include: SNTHERM (Jordan, 1991),

SNOWPACK (Bartelt and Lehning, 2002), CROCUS (Brun et al., 1989), SnowTran-3D (Liston

and Strum, 1996) and the Prairie Blowing Snow Model (Pomeroy et al., 1993). These models are

straightforward although not always simple to use and require computing power. Input variables

for these models are primarily gathered from local meteorological stations.

1.2.2 Regional Snow Models

Larger scale models are designed to cover a larger area and can become more complicated as

they deal with regional scale processes. These types of models range from spatial models of

catchment run-off processes with strong snow showing (UEB (Tarboton et al., 1995)) to large-

scale land surface models (CLASS (Verseghy, 2000), MESH(Pietroniro, et al., 2006)). These

larger models will typically incorporate other small models within them which allow them to

simulate a number of processes, often simultaneously. In order to run these models successfully,

a larger number of forcing variables are necessary, due to the larger number of processes being

simulated, as well as a larger amount of computing power. Input variables for these models are

usually gathered from (but are not limited to) a network of local meteorological stations and

depending on the size of the study area, can also be remotely sensed from satellites and aircrafts.

1.2.3 Global Climate Models

The largest scalar models are lumped into the grouping of global climate models (GCM). These

models cover very large areas, require a large amount of input variables, and take time and

computing power to run. Examples of these models include the CGCM1 (Flato et al., 2000), the

ECHAM models (Roeckner, 1992, 1996, 2003, Roesch and Roeckner, 2006) and the HadCM3

(Gordon et al., 2000) to name a few. Due to the scale of these models, input variables are

gathered remotely through satellite sensors.

Page 25: Multiscale modelling of snow depth over an agricultural

7

1.2.4 Model Challenges

The concept of scale is also very important as it determines the input requirements and the output

restrictions. Singh (1995) provides a brief outline into 25 different hydrological models. Of those

25 models, only 11 include snow processes (Singh, 1995). This demonstrates the problems that

many modellers face when attempting to design a model that can simulate the natural processes

that occur in the cryosphere during the snow accumulation and ablation season. In many cases,

the models do not consider every snow process and instead, focus on just a few. There is a

common belief that parameterization with accurate data and representative forcing data will

increase the model‟s accuracy. However, recent studies comparing many models clearly

demonstrate the inability of some models to accurately simulate the in situ conditions of snow

processes (Essery, et al., 2008, Rutter, et al., 2009). The SnowMIP (Model Inter-comparison

Project) was a comparison of atmospheric general circulation models, hydrologic snowmelt

models, numerical weather

prediction models, and detailed

avalanche and snow physics

models (Essery and Yang 2001).

Figure 1-2 demonstrates the

simulated results for snow water

equivalent (SWE) from the 24

models in the SnowMIP project.

The black lines are the model

runs from each model and the green

dots are the in situ measurements.

Figure 1-2: SnowMIP2 analysis from Boulder, CO. The solid black lines are

the model runs. The green dots are the in situ values. (from Rutter et al.,

2009)

Page 26: Multiscale modelling of snow depth over an agricultural

8

This project is a clear example of how difficult it can be to simulate in situ observations and how

site-specific models can be. One of the main problems with snow modelling is the fact that each

model has been designed to simulate a specific type of landscape and may not (or on the rare

occasion, may) be transferable from one location to another without serious error.

This thesis is concerned with the application of the Prairie Blowing Snow Model (PBSM)

(Pomeroy et al., 1993). The model is a physically-based collection of algorithms that work in a

deterministic lumped fashion. Distributed models need to be altered and usually coupled with

other models in order to work properly. It was first develolped in 1987 as a single column mass

and energy balance that calculates blowing snow transport and sublimation rates (Pomerory,

1988,1989) and later extended to include a snow cover mass balance for the case of two

dimensions (Pomeroy et al., 1993). The model has been updated several times to deal with issues

including varying fetch and land use (Pomeroy et al., 1993) and varying terrain (Pomeroy et al.,

1997; Essery et al., 1999). The current version that is used in this thesis is based on the 1997

version and is a modified, single column calculation with updated methods to calculate the inputs

and scales of the fluxes from a point to a landscape in an areal snow mass balance calculation. It

is located as a module within the 6.0 platform of the Cold Regions Hydrological Model (CRHM)

(Pomeroy, 2007). CRHM is a modular modelling environment with a library of hyrological

modules which can be selected and linked together to simulate the hydrological cycle in the

natural environment. In this thesis, CRHM is used as a diagnostic tool to examine the blowing

snow processes and seasonal changes in the accummulated snowpack. CRHM separates the

landscape into individual units termed „hydrological response units‟ (HRUs). HRUs are the

largest landscape units having definable hydrological characteristics, meaning they can be

described by unique sets of parameters, variables, and fluxes (Pomeroy et al., 2007b). In order

Page 27: Multiscale modelling of snow depth over an agricultural

9

for PBSM to work correctly it must be coupled to a snowmelt module. In this thesis it will be

coupled to the MSM (Minimal snowmelt model) model. Further details about PBSM and MSM

are found in the section devoted to the model descriptions.

1.3 Aims and Objectives

The aim of this thesis is to characterize snowpack changes throughout the 2008-2009 winter

season at Strawberry Creek, Ontario, and to consider the scaling effects of the model as it is

applied to various scales and in different formats with similar forcing data. To achieve this aim

three objectives are identified. The first of these objectives is to implement the PBSM and MSM

modules as a single model within the CRHM platform at the point scale. The second objective is

to implement the model over the entire agricultural field in a lumped format. The third objective

is to implement the model in a distributed format over the entire field.

a) Implementing PBSM and MSM at the point scale

The first objective is to implement the modules as a single model within the CRHM model

platform at the point scale and then assess the model performance at this scale. This scale is

defined as the area immediately surrounding the meteorological station (0.025km2).

b) Implementing the model over an agricultural field in a lumped format

The second objective is to scale up from the point scale and apply the model over the entire

agricultural field in a lumped format followed by assessing the model performance. This scale is

defined as the area contained within the property borders (0.406km2)

c) Implementing the model over an agricultural field in a distributed format

Page 28: Multiscale modelling of snow depth over an agricultural

10

The third objective in this thesis is to implement the model at the same field scale as objective 2

but in a distributed format. The field is subdivided into three HRUs which have been determined

based on vegetation height. The model performance is then assessed to determine the effects of

vegetation height across the entire field at that scale.

1.4 Thesis Layout

This thesis is a modelling study of snow distribution in a blowing snow model by taking both a

lumped and distributed approach. Chapter 1 covers the basics of hydrological modelling,

problems associated with modelling snow, as well as a quick overview of snow properties.

Chapter 2 provides a more in depth look at the physical properties of snow as well as energy and

mass balance systems. Chapter 3 covers the background information about the models and

modules used in this thesis. Chapter 4 outlines the study site and methodology of the project at

Strawberry Creek. Chapter 5 displays the field results, both meteorological and snow survey, as

well as the model results. Model performance and statistical analysis are also covered in this

chapter. Chapter 6 further discusses the model performance as well as some of the sources of

error that affect the model performance. Future considerations and concluding remarks round out

this chapter.

Page 29: Multiscale modelling of snow depth over an agricultural

11

Chapter 2.

Background – Snow accumulation and energy and mass exchanges through water

Seasonal snow falls and remains on the ground until later on in the season when it ablates. This

seasonal cycle of snowfall can be broken down into two phases; the accumulation phase and the

melting phase. The melt phase can be further broken down into phases of warming, ripening, and

output (Dingman, 2002).

The accumulation phase occurs when there is a general increase in water equivalent and a

decrease in temperature and net input of energy (Dingman, 2002). As snow falls through the

atmosphere it is subject to a variety of physical processes until it either becomes intercepted by

the vegetation canopy or is allowed to fall straight to the ground surface where it is further

subjected to physical and metamorphic processes over time. Once on the ground, these effecting

physical and metamorphic processes are dependent upon whether or not the snowpack is wet or

dry. A snowpack is considered „wet‟ when its temperature is at or above 0°C and a snowpack is

considered „cold‟ when its temperature is below 0°C. The accumulation phase is often dominated

by dry snow conditions. Under these conditions, the metamorphic processes depend on whether

the snowpack is isothermal or exhibits a temperature gradient (Male, 1980).

When the net energy within the snowpack is positive, the melt phase begins. The main processes

of melt are dependent upon climate and location. In temperate regions, radiation is the main

source of melt energy, followed by sensible and latent heat fluxes. Heat energy added by rain and

the ground can also play important roles during the melt phase in both temperate and arctic

environments.

Page 30: Multiscale modelling of snow depth over an agricultural

12

The water which is generated from surface snow melt, or input from rainfall, results in a wet

snowpack. Wet snow causes different metamorphic processes then dry snow because it causes

smaller snow particles to disappear and allows larger particles to grow and fuse together with

other particles. This results in a snowpack with less structural strength between the particles and

a higher density. Colbeck (1978) states that the following loss in surface area causes a drop in

capillary potential and may be the reason that a pack can be observed to release a lot of water

very quickly in the early stages of melt.

Within the melt phase, the warming phase is the period in which the snow temperature increases

to 0°C, the ripening phase is the period at which the melting snowpack can no longer hold any

more water, and the output phase is the period when the snowpack releases the water (Dingman,

2002). The differentiation between all three of these melt phases is not always straightforward. A

snowpack can undergo several stages before reaching a stage of isothermal energy due to the fact

that there are large energy gradients between the atmosphere, snow, and ground which can cause

the snow within the pack to melt and refreeze several times throughout the season. In many

cases, snow will accumulate and melt several times in one season, often re-accumulating upon

partially melted snowpacks. This can complicate the prediction of net snow water equivalent.

The following section will discuss some of the physical properties of snow as well as the energy

and mass balances of a snowpack.

2.1 Physical Processes of Snow

2.1.1 Precipitation

Snow forms in the atmosphere where cloud temperatures are below 0°C. The large diversity of

initial snow shapes is a result of the variety of temperature and humidity conditions possible

Page 31: Multiscale modelling of snow depth over an agricultural

13

during crystal formation. As the snow falls through the atmosphere, the changes in temperature

and humidity that the crystals encounter alters their form, compounded by collisions that further

modify their structure (Male 1980). Many factors affect the characteristics of snow from

formation to deposition. In the simplest of terms, precipitation can come in the form of rain or

snow, or as a mixture of rain and snow, depending on the meteorological conditions as

precipitation falls through the air. Further information on the various forms of precipitation,

specifically snow, can be found within The International Classification of Seasonal Snow on the

Ground (UNESCO/IHP, 2009).

2.1.2 Density

The density of snow is the mass per volume, usually specified in kilograms per cubic metre

(kg/m3). It is usually referred to as a ratio to the density of water (1000kg/m

3) and can vary

dramatically depending on temperature. The growth of individual snowflakes depends on

temperature and both the horizontal and vertical wind gradients. Predominately however, it is the

temperature that derives the overall size and shape of the snowflake (DeWalle and Rango, 2008).

Higher density snow usually results from warmer temperatures and/or higher winds whereas

lower density snow results from cooler temperatures and less wind. Warmer temperatures allow

snowflakes to rapidly decompose and coagulate with other snowflakes in the air or on the ground

surface. This increase in mass as well as the fact that the snowflake temperature is closer to the

melting point results in it having a higher density. Colder temperatures allow the snowflakes to

remain angular and faceted for a longer period of time in the air and on the ground surface. This

extended period of time results in a delayed coagulation of the snowflake with other flakes,

leading to lower density values. Higher wind speed can compact snow on the ground surface into

slab-like conditions, raising the density to a higher point then snow that has been left to settle at

Page 32: Multiscale modelling of snow depth over an agricultural

14

lower wind speeds. New snow will have a much smaller density value then older snow due to

snow settling. This is because older snow within a snowpack has been exposed to periods of

melting, re-freezing, and wind fluctuations. As soon as a snowflake comes in contact with the

ground surface it begins to change shape, losing its angular, faceted structure and developing into

a small cluster of grains. Due to melting and re-freezing, these small grain clusters can coagulate

with other grain clusters within the snowpack, developing a hard packed surface with a high

density value. Fresh snow density has been found to vary from 70 to 100 kg/m3 by Goodison et

al. (1981). Male (1980,) reported that fresh snow density can vary from 10 to 500 kg/m3. La

Chapelle (1961) related snow density to air temperature from data collected at the Alta

Avalanche Study Center and derived an equation for calculating the density of fresh snow which

can be seen in equation 2.1. By examining the works of Diamond and Lowry (1953) and Schmidt

and Gluns (1991), Hedstrom and Pomeroy (1998) developed another equation for calculating the

density of fresh snow, which can be seen in equation 2.2 below.

[2.1]

[2.2]

In both equations, ρs (fresh) is the fresh snow density (kg/m3), and Ta is the air temperature (°C).

Snow density increases exponentially with time, as wind, sublimation, gravity, and warm periods

change the internal structure of the snowpack. The maximum density of a snowpack ultimately

depends on its location, landcover, and other factors (Fassnacht, 2000).

The snow density can also be measured using both in situ measurements as well as ground-based

remotely sensed measurements. In situ measurements are gathered through the use of snow pit

and gravimetric samples. Remotely sensed measurements include microwave radar (Frequency-

Page 33: Multiscale modelling of snow depth over an agricultural

15

Modulated Continuous Waves and Ground Penetrating Radar) (Marshall et al., 2004), and

gamma ray attenuation.

2.1.3 Snow Depth

Snow depth is the accumulation of snow over a given period of time. The depth of snow at a

given area is a factor of the climatological conditions (temperature + precipitation), the

topography (slope + aspect), wind conditions (direction + velocity), as well as the land-use cover

and vegetation characteristics. Changes in snow depth can be the result of an increase in

accumulated precipitation after a snow storm, and can decrease due to melting. Furthermore,

snow depth varies in response to changes in snow density (e.g. decrease in volume without a

change to mass). For energy and mass balance purposes, it is more pertinent to measure a net

change in mass storage rather than just the depth of the snowpack. The snow cover depth can be

measured in situ or remotely sensed using an automated measuring device (e.g. MagnaProbe), a

snow ruler, graduated rod, aerial markers, and ultrasonic snow depth sensors (sonar).

2.1.4 Snow Water Equivalent

Snow water equivalent (SWE) is the equivalent depth of water of a snow covered area. Since a

uniform depth of 1mm of water spread over an area of 1m2 weighs 1kg, SWE is calculated from

the snow depth, ds, and the snow density, ρs, by the expression:

[2.3]

in which SWE is in mm when ds is in cm and ρs is in kg/m3 (Dingman, 2002). As the density of

the snowpack changes, so too will the SWE. An average density of 100kg/m3 is often assumed

for freshly fallen snow. This gives 1 unit of water equivalent for 10 units of snow depth. It is

Page 34: Multiscale modelling of snow depth over an agricultural

16

however, not appropriate to use a common conversion for all environments (Pomeroy and Gray,

1995). Once the amount of water equivalent is established, it becomes simple to determine how

much energy is needed to either melt or freeze this depth of water given the appropriate latent

heat values.

Term Remarks

Approximate

Range of θ

Graphic

Symbol

Dry

Usually T is below 0°C, but dry snow can occur at any

temperature up to 0°C. Dissaggregated snow grains have little

tendency to adhere to each other when pressed together, as in

making a snowball

0%

Moist T = 0°C. The water is not visible even at 10 x magnification.

When lightly crushed, the snow has a distinct tendency to

stick together.

<3%

Wet

T = 0°C. The water can be recognized at 10 x magnification

by its meniscus between adjacent snow grains, but water

cannot be pressed out by moderately squeezing the snow in

the hands. (Pendular regime)

3-8%

Very Wet

T = 0°C. The water can be pressed out by moderately

squeezing the snow in the hands, but there is an appreciable

amount of air confined within the pores. (Funicular regime)

8-15%

Slush T = 0°C. The snow is flooded with water and contains a

relatively small amount of air. >15%

Table 2-1. Classification scheme of SWE (Colbeck et al, 1990)

Table 2-1 shows that there are many different classification terminologies that can be assigned to

a certain percentage of liquid water content (θ). Much like the depth and density, the SWE can be

measured both in situ as well as by using a variety of remote sensors. In situ measurements

include gravimetric sampling and snow pillows. Remotely sensed measurements (satellites,

aircraft, etc) make use of algorithms and values of snow depth and density.

Page 35: Multiscale modelling of snow depth over an agricultural

17

2.2 Energy and Mass Balance Component Models for a Snowpack

The following section provides a basic theoretical background for many different mathematical

components that pertain to snow processes. It is important to have this background knowledge in

order to fully understand how the model arrives at its results as well as to better understand what

is naturally occurring in the environment.

2.2.1 Energy Balance components:

There are many different energy balance models that exist for snow but the differences between

these models are quite subtle. In order to completely understand the mathematical models

pertaining to snow it is necessary to view the snow in a finite condition. Figure 2-1 demonstrates

a typical block of snow and describes the energy balance associated with that element of snow of

surface area A and height hs.

Figure 2-1: A Block of Snow (Dingman, 2002)

Dingman (2002) provides a simple equation to explain this energy balance and can be seen

below:

[2.4]

Page 36: Multiscale modelling of snow depth over an agricultural

18

Where S (E L-2

T-1

) is the net rate of energy exchanges into the element by all processes over a

time period Δt (T), and ΔQ (E L-2

) is the change in heat energy absorbed by the snowpack during

Δt. Dingman expresses the measurable quantities of S, Δt, and ΔQ in terms of fundamental

physical dimensions of energy [E], length [L], and time [T]. The net rate of energy exchange, S

(W m-2

) can be expanded, as seen in Equation 2.5:

[2.5]

Where K (W m-2

) is the shortwave radiation input, L (W m-2

) is the long-wave radiation input, H

(W m-2

) is the sensible heat exchange with the atmosphere, LE (W m-2

) is the latent heat

exchange with the atmosphere, R (W m-2

) is the heat input from rain, and G (W m-2

) is the heat

input from the ground. Determining a complete energy balance for a given snowpack is

complicated. As will be discussed below, each term in Equation 2.5 can in turn be difficult to

acquire due to the constant fluctuation of shortwave radiation penetration depths and diurnal and

seasonal thermal variations within the snowpack.

2.2.2 Shortwave Radiation Input (K):

The net shortwave radiation input is described in Equation 2.6.

[2.6]

K↓ is the incoming shortwave radiation, K↑ is the outgoing shortwave radiation, and α is the

albedo. Albedo is the representation of the fraction of radiation reflected from the snow surface

and can vary from 0.04- 0.95, with new snow having the highest values and lesser values as the

snow ages (Oke, 1987). Since snow is highly reflective, even the shallowest of snow covers

alters the land surface albedo. Furthermore, the albedo of snow can vary diurnally with higher

Page 37: Multiscale modelling of snow depth over an agricultural

19

values occurring in the early morning and later in the evening when solar incidence angles are

low. At these lower angles, specular reflection dominates over diffuse reflection. Meteorological

conditions such as cloud cover, seasonality, and the physical state of the snowpack can all have

an influence on the albedo as well, allowing even more diurnal variation. The albedo for snow

also varies with wavelength where higher albedo is found with shorter wavelengths and lower

albedo with longer wavelengths.

Snow differs from other surfaces because it allows some transmission of shortwave radiation

throughout the snowpack. This means that at any depth, the shortwave radiation can be

transmitted (Ψ), reflected (α), or absorbed (ζ) according to the following equation (Oke, 1987):

[2.7]

It is interesting to note however, that this transmittance, absorption, or reflectance of shortwave

radiation is not universal throughout the vertical profile of the snowpack. The shortwave

radiation at the surface (K↓0) is much stronger than that found at any depth below. As the

shortwave radiation passes through the upper layers of the snow profile it begins to lose intensity

at an exponential rate.

Page 38: Multiscale modelling of snow depth over an agricultural

20

Figure 2-2: Exponential decay of solar radiation with depth for a typical snowpack and glacier. (Oke, 1987 after Geiger,

1965)

According to Beer’s Law:

[2.8]

the shortwave radiation reaching any depth z, is dependent on the shortwave radiation K↓z at

depth z, the base of natural logarithms e, and an extinction coefficient a (m-1

). The extinction

coefficient is dependent on the nature of the medium that is being transmitted through and the

wavelength of the radiation. As can be seen in the above diagram (Figure 2-2), snow has a larger

coefficient than ice, therefore snow has a much shallower penetration depth. This penetration

depth can be as great as 1m for snow and as much as 10m for ice (Oke, 1987).

The transmission of solar radiation within a snowpack can create problems when trying to

monitor or observe the radiation for energy balance purposes. Instruments mounted above the

snow surface may not only be measuring the reflectance from the surface but also some of the

reflectance from some of the subsurface layers. The resulting albedo calculation would therefore

be a volume and not a surface value.

Page 39: Multiscale modelling of snow depth over an agricultural

21

2.2.3 Long-wave Radiation Exchange (L*):

The net long-wave radiation exchange is described in Equation 2.9.

[2.9]

L↓ is the incoming long-wave radiation, L↑ is the outgoing long-wave radiation, εss is emissivity

of the snow surface, εat is the integrated effective emissivity of the atmosphere and canopy, and σ

is the Stefan-Boltzman constant (5.67 x 10-8

W m-2

K-4

). Tat is the absolute radiative temperature

of the atmosphere and canopy and Tss is the absolute temperature of the snow surface. Snow

(especially fresh snow) is almost a perfect radiator. The emissivity of snow is usually assumed to

be 1.0 but much like albedo, it too can vary with the age of the snow from 0.85 – 0.99 (Oke,

1987) as well as due to metamorphological processes (Grody, 1996). Even though the emissivity

can be relatively high, the surface temperature Tss is low, leading to L↑ being overall, quite small.

One of the major issues with implementing Equation 2.9 can be found in properly estimating

absolute values for εat and Tat, or equivalently, to estimate a value for L↓ under various

conditions of cloudiness and forest cover. Methods for estimating L↓ can be found in Dingman

(2002).

Oke (1987) presents a very interesting scenario for estimating L↑ when the snow surface is

melting (T0 = 0°C [273.2 K]). If we assume ε for snow to be a value of 1, and the surface

temperature to be at the freezing point, then the value of L↑ is constant at 316 W m-2

(substituting into equation 2.9). If this melting occurs under complete cloud cover, then the net

long-wave exchange (L*) between a fresh snowpack and the overcast sky becomes a function of

their respective temperatures due to the fact that clouds are also close to being full radiators. This

process can be seen in the following equation:

Page 40: Multiscale modelling of snow depth over an agricultural

22

[2.10]

where Tc4 is the cloud base temperature. If the cloud base temperature is warmer then the snow

surface temperature then L* will be positive. Under clear skies however, L

* is almost always

negative.

2.2.4 Sensible Heat Exchange with the Atmosphere (H):

When vertical temperature gradients and turbulent eddies exist above the surface, sensible heat is

transferred to the atmosphere. Equation 2.11 expresses the sensible heat exchange with the

atmosphere.

[2.11]

DH is the diffusivity of sensible heat, DM is the diffusivity of momentum in turbulent air, ca is the

heat capacity of air at a constant temperature (J kg-1

°C -1

), ρa is the density of air (kg m-3

), k is a

dimensionless constant (0.4), um is the velocity of wind (m s-1

) at height zm (m), zd is the zero-

plane displacement, z0 is the roughness height (m), Ts is the surface temperature (°C), and Tm is

the temperature of the measurement height (°C). This equation can further be adapted for snow if

we assume that both the wind speed, vm, and air temperature, Ta, are measured at the same

height, zm (typically 2m above the ground surface), that the diffusivity of sensible heat and

momentum are the same (DH =DM), and finally, that the zero-plane displacement, zd, is

negligibly small. With these assumptions Equation 2.11 becomes

[2.12]

Page 41: Multiscale modelling of snow depth over an agricultural

23

2.2.5 Latent Heat Exchange with the Atmosphere (LE):

When turbulent eddies and a vertical pressure gradient are both present, latent heat (water

vapour) is transferred to the atmosphere. Equation 2.13 expresses the latent heat exchange with

the atmosphere.

[2.13]

In addition to the terms that have already been defined above, DWV is the diffusivity of water

vapour in turbulent air, λv is the latent heat of vaporization (J kg-1

), P is the air pressure (mb), es

is the vapour pressure (Pa) at the surface and em is the vapour pressure (Pa) at height zm.

2.2.6 Heat Input from Rain (R):

There are two equations that apply to the heat input from rain. These two equations are

dependent on whether or not the snowpack is a „wet‟ snowpack at or above the freezing point

(Equation 2.14a) or if the snowpack is a „cold‟ snowpack below freezing (Equation 2.14b).

[2.14a]

[2.14b]

ρw is the density of water (kg m-3

), cw is the heat capacity of water (J kg-1

°C -1

), r is the rate of

rainfall (m s-1

), and Tr is the temperature of the rain (°C). The only additional term for the heat

input from rain in a snowpack below freezing is λf, which is the latent heat of fusion. If rain

percolates through the snowpack it becomes an additional heat source which can then melt the

surrounding snow and/or result in a possible addition to meltwater run-off.

Page 42: Multiscale modelling of snow depth over an agricultural

24

2.2.7 Heat Input from the Ground (G)

Summer storage of thermal energy results in temperatures in the soil profile to increase

downwards from the snowpack. The rate of heat that is conducted to the base of this snowpack is

described in Equation 2.15.

[2.15]

kG is the thermal conductivity of the soil (W m-1

°C-1

) and dT/dz is the vertical temperature

gradient in the soil. G becomes significant when the snowcover over top of the ground surface is

thin and there a heat exchange across the base of the snowpack (Oke, 1987). When the snowpack

is larger, it acts as a thermal insulator and the effect of G becomes rather unimportant (Oke and

Hannell, 1966). Dingman (2002) states that G is usually negligible during a snowmelt season,

but can be significant during the accumulation season.

2.2.8 A comparison of Four Different Energy Balance Equations:

The following section will examine several energy balance equations as they pertain to varying

types of snowpack and how they deal with the different types of energy. For a more in depth

comparison of snowmelt energy balances see Kuusisto (1986). The first of the four equations

was put forth by Dingman (2002) and described above as Equation 2.5. The second equation

comes from Marsh (1990) and is described below as Equation 2.16.

[2.16]

QM (W m-2

)is the available energy for melt (latent heat storage due to melting or freezing), Q*

(W m-2

) is the net all-wave radiation (K + L), QH (W m

-2) is the sensible heat flux, QE is the latent

Page 43: Multiscale modelling of snow depth over an agricultural

25

heat flux, QR (W m-2

) is the heat flux from rain, QG (W m-2

)is the heat flux from the ground at

the soil-snow interface, and dU/dt (KJ s-1

m-2

) is the rate of change of internal energy. The third

and fourth equations both come from Male (1980) and can be seen as Equations 2.17 and 2.18

respectively below.

[2.17]

[2.18]

In Equation 2.17, Qi (W m-2

) is the energy flux due to net radiation, sensible and latent heat

fluxes, and heat transfer from the lower layers of the snowpack. (ph)i is the energy transfer due to

precipitation, p (kg m-2

s-1

) with an associated specific enthalpy, h (KJ kg-1

). The second of

Male‟s equations, Equation 2.18, dU/dt (KJ s-1

m-2

) is the change in internal energy of the pack.

Male (1980, p.349) states that Equation 2.17 only requires measurements taken at or near the

upper surface and is therefore suitable for deeper snowpacks. Equation 2.18 describes the entire

thermal regime of the snowpack and can be practically used for shallow snowpacks of less than

40cm. When horizontal advection is substantial (i.e. not near open water, or patchy snow cover),

neither equation is considered valid.

Of all the equations presented above, Equation 2.17 is the simplest. Equations from Dingman

(Equation 2.5), Marsh (Equation 2.16), and Male (Equation 2.18) all consider the internal energy

of the snowpack but the latter two equations also include the melt energy. In Equation 2.16 the

melt energy remains in the form of energy but in Equation 2.18, the melt energy is actually the

melt-water from the bottom of the snowpack. Both of Male‟s equations focus primarily on the

melt phase but, as was previously mentioned, could be used for the accumulation phase so long

as the conditions of negligible horizontal advection are met. Dingman and Marsh‟s equations

Page 44: Multiscale modelling of snow depth over an agricultural

26

include both the accumulation and melt phases and are assumed to work for a snowpack of any

depth. For the sake of this thesis the above statements shall remain strictly theoretical but a quick

application of a common data set would help to better understand the differences among and uses

between each equation.

2.2.9 Mass Balance Components

Acquiring a proper mass balance for a given snowpack can be rather difficult due to the constant

phase changes occurring within the snowpack throughout the season. As was discussed in the

above sections, a snowpack can be influenced by a variety of different factors that can result in

diurnal and seasonal fluctuations in both the energy and mass balances.

Figure 2-3: Schematic diagram of the mass balance of a snowpack (after Oke, 1987)

A simple mass balance can be determined if we assume a snowpack where the top section is

exposed to the air interface and its lower section is either at the ground interface or is at a depth

of negligible water percolation (Figure 2-3), we can assume that the net change in mass storage

ΔS, is given by:

[2.19]

Page 45: Multiscale modelling of snow depth over an agricultural

27

Where p is the precipitation input from either snow or rain, E is the net turbulent exchange with

the atmosphere and can be in the form of condensation, sublimation (inputs) or evaporation and

sublimation (outputs), and Δr the net surface and subsurface horizontal exchange (surface snow

drifting and meltwater flow or subsurface meltwater throughflow) (Oke, 1987).

2.2.10 Sublimation, Evaporation, and Condensation of a Snowpack

As was mentioned above, water within a snowpack is susceptible to the physical processes of

sublimation, evaporation, and condensation. These processes are driven by the vapour pressure

gradient between the surface of the snowpack and the atmosphere. When the vapour pressure

above the snowpack is greater than the vapour pressure at the surface, condensation occurs.

When the vapour pressure above the surface is lower than the vapour pressure at the surface,

sublimation or evaporation occurs.

2.2.11 Canopy Interception

When a vegetated canopy is present, snow can become trapped by the branches and foliage of

the canopy, potentially inhibiting the trapped snow from reaching the snowpack on the ground

surface. While suspended in the canopy, snow is susceptible to the same physical processes as

snow on the ground surface and is not only significant to the overall energy and mass balance

equations but is also subject to its own energy and mass balances. Falling snow is not the only

method for which snow can become trapped within a canopy. Horizontally blown snow can also

become a significant source of trapped snow. Hedstrom and Pomeroy (1998) estimated the

snowfall canopy interception, I, as can be seen in Equation 2.20:

[2.20]

Page 46: Multiscale modelling of snow depth over an agricultural

28

where csuc is a dimensionless snow unloading coefficient found to be 0.697, I* is the maximum

snow load (kg m-2

), I0 is the snow load (kg m-2

), Rs is the snowfall for a unit of time, and Ccan is

the canopy coverage fraction.

The maximum snow load can be determined using Equation 2.21; where Sp is a tree species

coefficient, LAI is the leaf area index, and ρs(fresh) is the fresh snow density (kg m-3

).

[2.21]

For the purpose of this thesis, canopy interception is considered negligible. Even though the field

site is bordered by a small forest on one side and a thin treeline on another, the model has

difficulty simulating the effects from these canopies due to the exclusion of canopy interception

algorithms in the modules being used.

2.2.12 Blowing Snow

Blowing snow is a very important process as it is the main form of snow redistribution on the

ground and within the atmospheric boundary layer. This redistribution of snow can and will have

an effect on other processes such as the sensible and latent heat fluxes. It can be divided into two

parts: saltation and suspension. Sublimation also plays an integral role in blowing snow transport

and further information will be covered in the next few sections devoted to blowing snow and the

Prairie Blowing Snow Model.

2.3 Blowing Snow and the Winter Moisture Budget

The redistribution of snow by wind is a very important concept for any hydrological research

project located within the cryosphere. Wind redistributes snow by eroding it from areas of high

wind speed, such as ridge tops and windward slopes, and deposits it in areas of lower wind

Page 47: Multiscale modelling of snow depth over an agricultural

29

speed, such as the lees of ridge tops, vegetation stands, and topographic depressions (Benson and

Sturm, 1993; Pomeroy et al., 1993; Liston and Sturm, 1998; Sturm et al., 2001a; Liston et al.,

2002). When the prevailing winds constantly come from the same direction, a pattern of

heterogeneous snow drifts occur in the same location year after year. This redistribution can lead

to patterns of environmental conditions and ecosystem properties. Affected properties include net

solar radiation, chemical inputs, meltwater distribution, rock weathering, pedogenesis,

decomposition and mineralization, animal habitat, and vegetation (Hiemstra et al., 2002).

Snow is a loose surface, allowing saltation and suspension to act as the primary mechanisms for

distributing it by wind (Kobayashi, 1972; Male, 1980; Takeuchi, 1980; Schmidt, 1982, 1986;

Pomeroy, 1988; Pomeroy and Gray, 1990). Empirical and analytical relationship between wind

speed and blowing snow transport rate developed by these researchers can be found within their

snow-transport models and those developed by others (Tabler and others, 1990; Liston and

others, 1993a; Bintanja, 1998a,b; Sundsbø, 1997; Purves and others, 1998). Others have shown

that sublimation can be important, reducing the snowcover water balance by a significant

fraction under certain conditions (Goodison, 1981; Benson, 1982; Benson and Sturm, 1993;

Pomeroy and others, 1993, 1997). Modelling provides an alternative method for representing

winter wind-transported processes over various landscapes. It offers an opportunity to evaluate

the relative importance of driving variables, and producing scenarios based on alterations in

these variables. Model outputs can in turn be distributed across realistic terrain and vegetation

representations to produce landscape models that integrate key blowing snow processes

(Pomeroy et al., 1997, Liston and Sturm, 1998). The following section explains the winter

moisture budget and some of the associated problems that occur when attempting to acquire

information that can be used in a blowing-snow transportation model.

Page 48: Multiscale modelling of snow depth over an agricultural

30

2.3.1 The winter moisture budget

When the temperature falls consistently below freezing, the simplest way to describe the winter

moisture budget (in the absence of any net horizontal transport) is: snow-water-equivalent depth

(D) equals precipitation (P) minus sublimation (S) (Liston and Sturm, 2004, equation 2.22):

[2.22]

The use of this equation is based on a few assumptions including the fact that spatial variation in

erosion and deposition are negligible and that sublimation is easily modelled or measured.

2.3.1.1 Precipitation

Precipitation accumulates on the ground, building a snow cover that affects atmospheric and soil

temperatures by moderating conductive, sensible, and latent heat energy transfers between the

atmosphere, snowcover, and ground (Hinzman et al., 1998; Nelson et al., 1998; Liston, 1999).

Precipitation on the other hand has proven to be quite challenging to measure accurately. Solid

precipitation generally falls when it is windy and windy environments can pose significant

underestimation in solid precipitation amounts (Larson and Peck, 1974; Goodison et al., 1981;

Benson, 1982; Yang et al., 1998, 2000), however, in some extremely windy conditions, blowing

snow can be caught by precipitation gauges, causing an overestimation (Yang and Ohata, 2001).

Yang et al. (1998) stated that;

“Even the most complex and advanced shielded gauges, such as the WMO Solid Precipitation

Measurement Intercomparison Project‟s „octagonal vertical Double Fence Intercomparison Reference‟

(DFIR), are unable to measure the true precipitation (liquid or solid) and require corrections for wind

speed.”

These corrections are not that large but are necessary. The standard unshielded 8-inch

precipitation gauge used by the National Weather Service (NWS) has an undercatch of nearly

Page 49: Multiscale modelling of snow depth over an agricultural

31

75% when exposed to windy environments (Goodison et al., 1981; Benson, 1982; Yang et al.,

1998, 2000). Others have shown that another form of gauge, the shielded Wyoming gauge, has

an undercatch of 6% for snowfall in non-windy conditions and 55% for blowing snow (Rechard

and Larson, 1971; Benson, 1982; Goodison et al.,1998; Yang et al., 2000). This is not only

important at the local scale but also at the regional and global scale. Precipitation gauge networks

are a crucial component to larger scale projects but there are very few networks that contain a

large amount of gauges. This means that one gauge is responsible for covering a very large area

which would induce a significant amount of potential error. There is some advocacy to increase

the number of monitoring stations that record precipitation across the Arctic but the reality of is

that this number is actually decreasing rather than increasing (Shiklomanov et al., 2002).

2.3.1.2 Snow depth measurements

In addition to the problems associated with precipitation gauges, there are also issues that arise

when attempting to record snow depth. The most accurate method of acquiring snow depth

measurements is to measure the amount of snow on the ground. Direct measurement using snow

courses (Pomeroy and Gray, 1995) and watershed surveys (Elder et al., 1991) can be effective

but time consuming, not to mention costly. It can even be very dangerous in certain terrain (e.g.

in avalanche zones). Liston and Sturm have found that due to the occurrence of blowing snow

and drifting snow, the snow depth and SWE distribution can be highly heterogeneous (Liston

and Sturm, 2002; Sturm and Liston, 2003). Both topography and vegetation play an important

role in this distribution and in many cases; require the use of models to help finalize the snow

depth amounts. Snow distribution is easier to estimate in flat terrain where there are no drift traps

present. It is assumed that the amount of snow being removed from one location due to blowing

snow and drifting will be countered by the same amount being brought in by the same

Page 50: Multiscale modelling of snow depth over an agricultural

32

mechanisms. Under these conditions, Equation 1 is valid. Otherwise, the use of blowing snow

models are important in acquiring values for snow depth in any area, large or small.

2.3.1.3 Sublimation

If snow is not caught in drift traps, and if the blowing snow event lasts long enough and the air

remains unsaturated, then particles can sublimate away. Tabler (1975) introduced the concept of

a particle transport distance – the distance an average-sized snow/ice grain can travel before it

completely sublimates away. Tabler (1975) found that for Wyoming, the average transport

distance was 3km. Following Tabler‟s methods, Benson (1982) found that this distance decreases

to 2-3km for the Arctic. There appears to be two separate schools of thought when it comes to

snow sublimation. One school of thought states that the highest rates of sublimation occur when

the snow particles are moving. Schmidt (1982) states that when the particles are in the wind

stream, their high surface-area to mass ratios and high ventilation velocities are responsible for

the high rates of sublimation. This research also demonstrated that these rates are as much as two

orders of magnitude higher than those for static snow under similar conditions (Schmidt, 1982).

The dependence of blowing snow sublimation rates on air temperature, humidity deficit, wind

speed, and particle size distribution, can be seen in the works of Tabler (1975), Lee (1975),

Schmidt (1982), and Pomeroy and Gray (1995). Schmidt‟s work has led to the creation of

blowing snow models such as the Prairie Blowing Snow Model (Pomeroy et al., 1993) and

SnowTran-3D (Liston and Sturm, 1998) to name a few. These models base their estimation of

sublimation as a direct result of wind transportation and do not include static snow sublimation

rates.

Page 51: Multiscale modelling of snow depth over an agricultural

33

Contrary to this school of thought, other studies have shown that blowing-snow sublimation is

not as important as sublimation from a static snow surface (Mann et al., 2000; Déry and Yau,

2001, 2002; King et al., 2001). In these studies, it is suggested that the atmospheric boundary

layer (ABL) quickly becomes saturated under blowing snow events and is therefore unable to

promote further ABL sublimation. Déry and Yau (2002) have found that static snow sublimation

rates were 5-10 times higher than those found in blowing snow conditions.

Liston and Sturm (2004) suggest that perhaps there may indeed be two different sublimation

regimes, with the interaction between the ABL and the overlying air mass determining which

regime dominates. They suggest both a “humid” regime and a “dry” regime. A “humid” regime

exists when the ABL is unable to mix with the overlying air mass and that the air is already

saturated. In this case static-snow sublimation is the dominate regime. Contrary to this, a “dry”

regime exists when the ABL is able to continuously mix with an overlying dry air mass, wicking

away any moisture from the near surface layer of air and from snow particles. In this case,

blowing-snow sublimation is the dominant regime (Liston and Sturm, 2004).

2.4 Summary

This chapter is focussed on providing background information about the physical properties of

snow and energy and mass exchanges through water. Information, importance, and data

collection of the physical processes of snow including precipitation, density, snow depth, and

snow water equivalent are all discussed. Energy and mass balance component models are also

discussed. Within the energy balance components there is a focus on the radiation budget as it

pertains to snow as well as a comparison of several different energy balance equations from

multiple sources. Mass balance components include changes to the snowpack through

Page 52: Multiscale modelling of snow depth over an agricultural

34

sublimation, evaporation, and condensation, canopy interception, and blowing snow. Blowing

snow is a very important component to this thesis and is further discussed as it pertains to the

winter moisture budget. Multiple trains of thought exist as to what is actually occurring in the

atmospheric boundary layer and these trains of thought are discussed.

Page 53: Multiscale modelling of snow depth over an agricultural

35

Chapter 3

The Models

3.1 The Prairie Blowing Snow Model

The prairie blowing snow model (PBSM) is an ensemble of physically based blowing snow

transport and sublimation algorithms which calculates the distribution of seasonal snowfall in

agricultural areas of the Canadian prairies. Snow transport is divided into two parts: saltation –

the movement of particles in a „skipping‟ action just above the snow surface, and suspension –

the movement of particles suspended by turbulence in the atmospheric layer extending to the

surface boundary-layer. Sublimation rates are calculated for a column of saltating and suspended

blowing snow extending to the top of the boundary layer (Pomeroy et al., 1993). The model was

designed to be able to predict: 1) mean vertical profiles of snow transport flux; 2) the effect of

grain stubble height on blowing snow transport and sublimation losses from snow cover; 3) the

effect of land use and unobstructed fetch distance on blowing snow transport and sublimation

losses from snowcovers. The process algorithms used that make up the model make use of

readily available meteorological information such as wind speed, temperature, humidity

measured at a single height, occurrence of blowing snow and land use information.

3.1.1 Saltation

The occurrence of high winds over a loose surface can allow particles to become induced into

several types of aeolian movement: saltation, suspension and, in some instances, surface 'creep'

(Kind, 1990). Saltation is the process by which particles are transported horizontally by bouncing

along the surface and is usually initiated by the collision of one particle into another (Bagnold,

1941) (Fig. 3-1). The distance travelled by a saltating particle is determined in part by the mean

flow of the fluid, the particle size and surface conditions (Pomeroy and Gray, 1995).

Page 54: Multiscale modelling of snow depth over an agricultural

36

Figure 3-1: Diagram showing the processes of saltation and suspension of sediments into the atmosphere (adapted from

Greeley and Iversen, 1985

Pomeroy and Gray (1990) found that the expression for the vertically integrated saltation

transport rate, Qsalt (kg m-1

s-1

), used in the PBSM is as follows

[3.1]

where Csalt is an empirically derived constant found to equal to 0.68 ms-1

, ρ is the air density (kg

m-3

), g is the acceleration due to gravity (m s-2

), u* is the atmospheric friction velocity (m s

-1) and

the subscripts n and t refer to the friction velocity (shear stress) applied to the non-erodible

surface elements (e.g. rocks and bushes that are protruding through the snow) and to the snow

surface, respectively, at the transportation threshold. This friction velocity is dependent on the

wind speed and the surface roughness and can be expressed as follows

[3.2]

where k is von Kàrmàn‟s constant (0.4), u(z) is the wind speed at height z and z0 is the

aerodynamic roughness height. Further explanations and examples of the threshold velocities of

non-erodible and snow surfaces can be found in Pomeroy et al. (1993). McEwen (1993) found

Page 55: Multiscale modelling of snow depth over an agricultural

37

that, close to the surface, the wind profile is modified because some of the sheer stress is carried

by the saltating or suspended particles. As well, the apparent roughness length is also modified.

Pomeroy (1988) found that the roughness height for blowing snow over complete snowcovers

can be expressed as follows

[3.3]

in which Cz is a dimensionless coefficient found equal to 0.1203 for prairie snowcovers, Cst is a

coefficient equal to 0.5m, Nst is the number of vegetation elements per unit area and Ast is the

exposed silhouette area of a single typical vegetation element which can be expressed as the

product of the stalk diameter and the exposed stalk height (total stalk height minus snow depth).

As an example, a typical value of Nst for a wheat field in Saskatchewan is 320 stalks m-2

(Pomeroy et al., 1993). Under high threshold conditions an efficient, particle-surface interaction

enhances saltation transport at high wind speeds but inhibits transport at low wind speeds. This

leads to an approximately linear relationship between wind speed and the saltating transport rate.

3.1.2 Suspension

Suspension of snow can only occur once particles that are already saltating have reached a

terminal velocity that is balanced by the upward-moving winds of the atmospheric boundary

layer due to turbulence. Calculating the rate of transportation for suspended snow can be found

by integrating the mass flux over the depth of flow, which can be classified as the area between

the top of the saltation layer to the top of the surface-boundary layer. Schmidt (1982) defined the

mass flux as the mass concentration multiplied by the mean downwind particle velocity. On

average this velocity is equal to the velocity of the parcel of air. From the equation for u*, the

transportation rate of suspension, Qsusp (kg m-1

s-1

), is

Page 56: Multiscale modelling of snow depth over an agricultural

38

[3.4]

where h* and zb are the lower and upper boundaries of the suspension layer, respectively, n(z) is

the mass concentration of suspended snow (kg m-3

) at height z (m). Further information and

calculations of n(z) can be found in Pomeroy and Male (1992) and Pomeroy and Gray (1990).

The upper boundary height, zb, is assumed to be dependent on the fetch over uniform, mobile

snow, however in the PBSM code there is a 5m upper limit that is often imposed for snow

transport calculations. Pomeroy (1998) suggested that snow particles that have been lifted to

heights above 5m are unlikely to settle back to the surface before sublimation removes much of

their mass. Therefore, for „surface hydrological purposes‟, only the mass flux below 5m heights

are considered when calculating the snow transport rates. In order for flow to fully develop, a

minimum fetch distance of 300m is required. Knowing this, the modelled equation for zb can be

written as

[3.5]

where x is the downwind distance from the beginning of the fetch over which the surface

roughness and the aerodynamic roughness, z0, can be considered uniform. This equation is a

modified version that now incorporates the value of 0.3, which was taken from Takeuchi (1980)

as the value for the top of a plume of diffusing particles at an elapsed time, zp(t). Following

Pomeroy and Male (1992), the estimations for the value of h* can be expressed as

[3.6]

where ch is a coefficient found equal to 0.08436 m-0.27

s1.27

.

Suspension transport increases as a power function of wind speed (Pomeroy, 1988). While the

presence of exposed crop stubble increases the wind speed required to attain particle suspension,

Page 57: Multiscale modelling of snow depth over an agricultural

39

an increase in turbulence due to this exposed stubble also increases the suspended transport rate

for a given wind speed above this limit.

3.1.3 Sublimation

Sublimation also plays a key role in blowing snow transportation rates. It can act as a significant

source of water vapour and as a sink for sensible heat in the air. It may also lead to a decrease in

size (radius) and mass of snow and ice particles, resulting in a reduction in their drag

coefficients. This means that in real terms, the problem of blowing snow can incorporate a

multitude of particle sizes and therefore should be dealt with in such a manner. The model

however, treats the problem with a mean particle size from a certain distribution of particles in a

steady-state atmosphere. Pomeroy and Gray (1994, 1995) have argued that, in the Canadian

Prairies, up to 75% of the annual snowfall over a one-kilometre fallow field is eroded by wind,

and that, typically, up to half of this amount is sublimated before deposition occurs. According to

Schmidt (1972, 1991), the rate of sublimation of a snow particle may be modelled by a balance

between radiative energy exchange, convective heat transfer to the snow particle, turbulent

transfer of water vapour from the snow particle and the consumption of heat by sublimation. If

we assume thermodynamic equilibrium, the rate of change of mass with respect to time can be

expressed as follows

[3.7]

(for more information on the derivation of this equation, see Schmidt 1991) where r is the radius

of a snow particle containing mass, m; σ (dimensionless and negative) is the water vapour deficit

with respect to ice (e-ei)/ei, where ei is the saturation vapour pressure (svp) over ice, and

approximately equal to relative humidity minus one (RH-1); Qr is the radiant energy received by

Page 58: Multiscale modelling of snow depth over an agricultural

40

the particle; λs is the latent heat of sublimation (2.838 x 106 J kg

-1); M is the molecular weight of

water (18.01 kg kmol-1

); R is the universal gas constant (8313 J kmol-1

K-1

); D is the diffusivity of

water (m2s

-1); Ta is the ambient air temperature; ρsat is the saturation density of water vapour (kg

m-3

); Ka is the thermal conductivity of air (W m-1

K-1

); and Nu and Sh are the Nusselt and

Sherwood numbers, respectively. The Nusselt number is a dimensionless number that represents

the ratio of convective to conductive heat transfer. The Sherwood number is also a dimensionless

number (also referred to as the mass transfer Nusselt number) which represents the ratio of

convective to diffusive mass transport. Lee (1975) found that adjacent to a blowing snow particle

in a turbulent atmosphere Nu is identical to Sh (Nu≡Sh). Both terms are dependent on Reynolds

number, Re(r,z) as

[3.8]

in which,

[3.9]

where Vr is the ventilation velocity and v is the kinematic viscosity of air (1.88 x 10-5

s-1

). In the

suspension layer, one could ignore the effects of turbulence on Vr (Schmidt, 1982b; Pomeroy and

Male, 1986) resulting in Vr being equal to the terminal velocities (ws) of the particles expressed

as

[3.10]

In the saltation layer, the ventilation velocity is divided into vertical (Csalt u*) and horizontal

components (2.3ut*)(Pomeroy and Gray, 1990). We already know the value of Csalt from

Equation 2 as 0.68, thus Vr can be written as

[3.11]

Page 59: Multiscale modelling of snow depth over an agricultural

41

In terms of order of magnitude, Pomeroy (1988) argues that “the speed of the saltating particles

relative to the wind speed is equal to the speed of the saltating particles relative to the surface”

The rate of change in mass of a particle requires the specification of the undersaturation of water

vapour, σ(=RH-1) given by

[3.12]

where σ2 is the undersaturation at z = 2m when there is a decrease in relative humidity with

increasing height during blowing snow. When the opposite occurs, Déry and Taylor (1996) have

shown that the undersaturation can be given by

[3.13]

Taken from Déry and Taylor (2006), the sublimation rate (negative) per unit area of snowcover,

qsubl (kg m-2

s-1

), can be determined by integrating the sublimation loss rate coefficient [csubl(z)]

for the entire column of blowing snow extending from the snow surface to the top of the

suspension layer, multiplied by the frequency of a certain particle size and its mass, over the

particle spectrum and over height. Giving us

[3.14]

where dr is the particle size increment (m) and N is the total number of particles per unit volume.

Further information on the sublimation loss coefficient, frequency distribution, and particle

radius can be found in Pomeroy et al. (1993) and Déry and Taylor (1996).

3.1.4 Erosion and Deposition

The PBSM is also able to calculate the rate of surface erosion and deposition at distance x,

downwind of the leading edge of a fetch qv(x,0). This rate is established by the net mass of snow

entering or leaving a controlled volume of atmosphere, which extends from the snow surface to

Page 60: Multiscale modelling of snow depth over an agricultural

42

the top of the surface boundary-layer for blowing snow, in x and z directions and through

sublimation occurring within this volume as well (Fig 3-2).

This erosion/deposition rate can be expressed as follows

[3.15]

in which the vertical flux at the top of the controlled volume, qv(x, zb), is equal to the negative of

the snowfall rate. In order for flow to fully develop, the surface erosion rate must equal the

sublimation rate minus the snowfall rate and will develop under invariant atmospheric and

surface conditions so long as there is an adequate fetch of mobile snow. The amount of fetch will

vary depending on land use and land cover. Pomeroy (1988) suggests a distance of 500m for full

development of flow to a height of 5m. Deposition on the other hand, occurs when the surface

roughness is great enough to, or there are topographic depressions that, decrease the wind speed

and the saltation and the suspension rates of transport, or when the snowfall is greater than the

surface erosion rate.

Figure 3-2. Cross sectional view of controlled volume of atmosphere extending from the snow surface to the top of the

boundary-layer for blowing snow (Pomeroy et al., 1993)

Within CRHM, the snow mass balance is dealt with slightly differently than the original model.

The snow mass balance for an HRU is due to the distribution and divergence of blowing snow

Page 61: Multiscale modelling of snow depth over an agricultural

43

fluxes in and around said HRU. With a fetch, x (m), the following expression for mass balance

can be drawn over an HRU,

[3.16]

where dSWE/dt is the surface snow accumulation rate (kg m-2

s-1

), P is the snowfall rate (kg m-2

s-

1), p is the probability of blowing snow occurrence within the HRU, F is the downwind transport

rate (kg m-2

s-1

), E is the snow surface sublimation rate (kg m-2

s-1

), EB is the blowing snow

sublimation rate (kg m-2

s-1

), and M is the melt rate (kg m-2

s-1

). Proper application of the blowing

snow algorithm requires every term to be solved for. This is why PBSM is coupled with MSM in

this thesis. Nipher snowfall gauge undercatch is corrected for in the PBSM module and SWE

accumulation is calculated as a residual of snow transport, snowfall, and sublimation rates. The

other terms are found when linked to a snowmelt module. Interval data for temperature, relative

humidity, and wind speed are used to calculate the transport and sublimation of blowing snow.

Using a unique parameterization scheme of employing cascading HRUs, PBSM, as implemented

within CRHM, enables the parameterization of the flow of snow transport from HRU to HRU.

Snow is redistributed amongst HRUs by basing it upon snow transport calculations, HRU size,

vegetation characteristics, and a distribution factor, Dp. By implying the distribution factor,

PBSM allocates blowing snow transport from areas that are aerodynamically smoother (windier)

to areas that are aerodynamically rougher (calmer) within a basin. The distribution factor value

ranges from 0 to 10, 10 being the roughest (Fang and Pomeroy, 2009).

Page 62: Multiscale modelling of snow depth over an agricultural

44

3.2 Minimal Snowmelt Model

In order for PBSM to work properly it needs to be coupled with a snow melt model. CRHM has

a few different melt modules to choose from but the minimal snowmelt model (MSM) was

chosen for this thesis based on its simplicity and its ease of use using the available collected

meteorological data.

The MSM module is based from Essery and Etchevers, 2004. It is a simple snow model with

only three adjustable parameters (fresh snow albedo, albedo decay rate for melting snow, and

surface roughness length). It performs the energy and mass balances with a focus on the surface

of the snowpack. Changes in the internal snowpack energy, energy from precipitation, and

energy from the ground are all neglected. It is driven by average inputs of incoming shortwave

(K↓) and longwave (L↓) radiation respectively, air temperature, specific humidity (Q1), wind

speed (u), and snowfall rate (Rs). Surface pressure (Ps) and the measurement height (z1) both

must be specified as well.

MSM follows common procedures used in snow models and land surface schemes to calculate

radiant and turbulent exchanges. Surface fluxes of sensible heat (H) and latent heat (LE)

exchanges and can be seen as follows,

[3.17]

and

[3.18]

Page 63: Multiscale modelling of snow depth over an agricultural

45

where ρ and Cp are the density and heat capacity of air respectively, Qsat (Ts,Ps) is the saturation

humidity at the snow surface temperature (Ts) and pressure (Ps), CH is a surface exchange

coefficient. Net radiation (R) absorbed by the snow surface is calculated by

[3.19]

where α is the snow albedo and σ is the Stefan-Boltzmann constant.

MSM also deals with albedo decay and atmospheric stability. Further information about the

above as well as the derivation for CH can be seen in Essery and Etchevers, 2004.

Due to the neglecting of heat fluxes within the snowpack, heat advection from precipitation, and

the substrate, the energy balance becomes

[3.20]

where λs and λf are the latent heats of sublimation and fusion respectively. M is the melt energy

(W m-2

). The mass balance is then calculated following

[3.21]

3.3 Summary

This chapter provides a brief synopsis of the Prairie Blowing Snow Model and the Minimal

Snowmelt Model. It highlights the key equations and derivations from the models as they help

lead to a better understanding of the experiments of this thesis. PBSM is a collection of

physically based blowing snow transport and sublimation algorithms and is used to calculate

transportation rates in the forms of saltation and suspension as well as sublimation rates. The

model also calculates erosional and depositional rates with the aid of the snowmelt model MSM.

Page 64: Multiscale modelling of snow depth over an agricultural

46

MSM is a simplified snowmelt model that ignores internal energy exchanges within the

snowpack, energy from precipitation, and energy from the ground. This simplified model is

easily coupled with PBSM using the collected met data and provides the missing variables to

complete a mass balance equation that includes blowing snow fluxes. Both PBSM and MSM are

modules found within the 6.0 platform of CRHM which is the main modelling platform used in

this thesis.

Page 65: Multiscale modelling of snow depth over an agricultural

47

Chapter 4.

Methods

4.1 Study Site

The study site for this thesis project is located on a small agricultural farm within the Strawberry

Creek watershed (0.4054 km2) which is about 15km northeast of Waterloo, Ontario, Canada

Figure 4-1: Location of Strawberry Creek field site

Page 66: Multiscale modelling of snow depth over an agricultural

48

(43° 33‟ 1”N, 80° 23‟ 24”W). This domain can be seen Fig 4-1 and Fig 4-2. Strawberry Creek is

a first order stream that flows into Hopewell Creek and eventually flows into Lake Erie through

the Grand River. The Grand River is the largest catchment basin in southwestern Ontario with a

total flow collection of approximately 6800km2 (Chapman and Putnam, 1984). Figure 4-3

illustrates the Grand River catchment.

The physiography of Strawberry Creek was shaped by late Quaternary processes and is located

in an area referred to as the Guelph Drumlin field (Chapman and Putnam, 1984). Glacial till is

the predominant sub-surface physiology in this area and is assumed to be comprised of

approximately 50% sand, 35% silt, and 15% clay (Chapman and Putnam, 1984). The landscape

is predominantly gently rolling till plains which primarily consist of agricultural fields and very

little pasture. The dominant field crops in the area consist of hay, corn (for grain), soybeans, corn

(for silage), and winter wheat. Several eskers and moraines are located in this region as well

(Karow, 1968). Further information about the physiography, soil structure, vegetation, and

human development of this area can be seen in Chapman and Putnam (1984), Harris (1999), and

Karrow (1974, 1993). The purpose of this subsection is to provide a simple overview of the study

site as it pertains to the over-land flow of wind-transported snow.

The farm (0.4km2) is located just outside the small town of Maryhill. It consists of three separate

vegetated fields (N, NE, SW), the creek running SE through the centre of the farm, and a farm

house and barn located in the centre of the land plot. The N and SW fields are separated from the

NE field by the creek and the N and SW fields are separated by the farmhouse and its driveway.

The N and NE fields were tilled in the late fall, and are therefore unvegetated in the winter and

have small rows of raised earth ridges (~5cm) as a result from the tiller. The SW field has a

winter grass crop with an average height of 10cm.

Page 67: Multiscale modelling of snow depth over an agricultural

49

a)

c)

e)

g)

b)

d)

f)

h)

Fig 4-2: A 360 view of the field site from eight cardinal directions. a) Northwest, b) North, c) Northeast, d) East, e)

Southeast, f) South, g) Southwest, h) West.

Page 68: Multiscale modelling of snow depth over an agricultural

50

The topography is gently sloping towards the creek with roughly a 9m difference in elevation

between the SW and NE corners to the creek bed (336-345m). There is also a gentle hill located

in the centre of the SW field. The creek is narrow, roughly 1.5m deep and 2m wide under bank-

full conditions and has a general U shape. Dredging of the creek in the past has increased the

Figure 4-3: Grand River watershed (GRCA, 2008)

Page 69: Multiscale modelling of snow depth over an agricultural

51

height and steepness of the banks and has also altered the riparian zone on either side.

This study site was chosen for several reasons:

1) It is a good representation of prairie-like conditions in southern Ontario. This is important as

the focus of this thesis is to test the model, which largely consists of the Prairie Blowing Snow

module, in a small, local catchment under different climatic conditions.

2) It is located in an area that is in closer proximity to the University of Waterloo so that it is

easy to get to for the purpose of performing multiple snow courses and data collection over the

entire snow season.

3) A good working relationship between the land owners, farmers and the Universities of

Waterloo and Wilfrid Laurier is already present, making accessibility straightforward.

4.2 Digital Imagery and Data

Digital data coverage for Strawberry Creek included orthoimagery, contour mapping (2m

resolution), road and water networks, and township field boundary maps. All of the data were

collected from the Grand River Conservation Authority‟s Grand River Information Network

(GRIN) (GRCA, 2008), and was used in ArcGIS 9/ArcView version 9.3 Geographical

Information System (GIS) (Environmental Systems Research Institute, Inc. [ESRI], Redlands,

CA). The orthoimagery was collected as a series of digital aerial photography in the spring of

2006 as a larger project named the South West Ontario Orthophotography Project (SWOOP).

Tiles 548_4822 and 548_4820 were used in this project and have a ground pixel size of 30cm

with an estimated positional accuracy of +/-1 metre.

Page 70: Multiscale modelling of snow depth over an agricultural

52

4.3 Meteorological data

Half-hourly measurements of windspeed and direction, temperature, relative humidity,

precipitation (rain), snow depth, solar radiation, and air pressure, were taken from a 3m

meteorological tower which was located in the centre of the SW field (as seen in Fig. 1-1). Data

was stored on a Campbell Scientific CR1000 data logger and was uploaded using Campbell

Scientific‟s LoggerNet software package. Error free solid precipitation (snow) is difficult to

obtain (see section on winter moisture budget). Therefore, in order to achieve a value of this

precipitation, snow depth was recorded by the sonic snow depth sensor for time series analysis.

If any significant increase in depth occurred, it was recorded as an increase in precipitation.

Snowfall amounts were generated by examining the difference between precipitation values on a

daily basis and them added together to create a daily snowfall amount. All of the collected

meteorological data was assumed to be consistent across the entire study domain. Table 4-1

demonstrates all of the sensors used on the met tower and their purposes. The meteorological

observations were used to drive the model for the period of 21 November 2008 to 12 February

2009.

Meteorological Condition Instrument

Wind Speed and Direction 05103-10 RM Young wind monitor CSC Spec

Temperature and Relative Humidity HC-S3-XT relative humidity and Air temp probe

Solar Radiation CNR1 KIPP and ZONEN net radiometer sensor

Precipitation (rain) TE525M Texas Instrument tipping bucket

Barometric Pressure 61205V RM Young Barometric pressure sensor

Snow Depth SR50A Sonic ranger 50 KHz module

Table 4-1: List of selected meteorological instrumentation

Page 71: Multiscale modelling of snow depth over an agricultural

53

Over the entire winter season there were two significant melt periods where the snow cover

completely melted away except in areas of significant drifting. The first occurred in late

December, the second occurred in early February. Both melt periods were the result of a

combination of warmer temperatures and rain events. Due to instrument malfunction of the

SR50a sonic sensor, data from after the second severe thaw period (Feb. 12/09) was discarded.

This truncated the entire data set to only 83 days.

4.3.1 Ultrasonic snow depth sensors vs. traditional snow measurement techniques

Ultrasonic snow depth sensors (USDS) have been around since the early 1980s (Goodison et al.,

1984). They send out a sound pulse (50 kHz) and measure the time it takes the pulse to reach the

ground and return to the sensor. The pulse is projected downwards over a cone of 30 and for

accurate measurement it is imperative that there are no interferences between the sensor and the

ground surface. The time for the pulse to return to the transducer in the sensor is adjusted for the

Figure 4-4: View of meteorological tower looking south towards the town of Maryhill, Ont.

Page 72: Multiscale modelling of snow depth over an agricultural

54

speed of sound in air based on measured air temperature, and the timing is converted to a

distance via internal algorithms. To adjust the speed of sound in air (Vsound) in meters per second

for the ambient air temperature (Ta) in kelvins, the following equation is applied:

[4.1]

The distance the sound pulse travels decreases as snow accumulates on the ground, thus reducing

the time for the pulse to return to the sensor.

Traditional snow measurements consist of gauge precipitation, snowfall, and snow depth. Gauge

precipitation is defined as the liquid equivalent obtained by a nonrecording/recording, standard

precipitation gauge. As is mentioned in the winter moisture budget section of this thesis, gauge

precipitation is greatly affected by wind speed and direction and can lead to erroneous values

obtained for evaluation. Snowfall is defined as the maximum accumulation of new snow since

the last observation and is customarily measured manually with a ruler. Snow depth is defined as

the total depth of snow on the ground at the time of observation and indicates both old and new

snow on undisturbed surfaces. The measurement of snow depth may be the average of several

total depth measurement to obtain a representative sample (NWS, 1996).

USDS measurements are done with respect to manual measurements which are assumed to be

„ground truth‟. Uniformity and consistency in manual measurements are difficult to attain with

any data collector. There is an inherent uncertainty in manual measurements due to the fact that

snow melts, settles, blows, and drifts and does not accumulate uniformly on the ground.

Depending on time of day, the frequency, the measurement surface, the extent of nonuniformity

in snow accumulation, and overall care and detail of the individual observers, some variation in

Page 73: Multiscale modelling of snow depth over an agricultural

55

manual observation is to be expected. According to Ryan et al. (2008), there are no known

studies that have attempted to quantify this uncertainty in terms of measurement error, but they

estimate it to be in the magnitude of ± 25%.

The choice to use a Campbell Scientific SR50a sonic sensor to derive solid precipitation values

over traditional measurement techniques was made in part for the ease of obtaining an automated

daily snow depth measurement based on hourly readings which can be personally monitored.

Obtaining a manual snow depth measurement on a daily basis at the field site for the entire

season was feasibly impossible. A daily measurement of snow on the ground was manually

recorded every morning at the Waterloo-Wellington International Airport located in Breslau, 10

km to the southwest but due to any number of potential issues mentioned above, this data was

not used as an input in any model runs for this thesis. A comparison between the SR50a data

from Strawberry Creek and the manual snow depth data from the airport is seen in Fig. 4-5

below. The average seasonal difference between the two data sets was approximately 8.0 cm.

Figure 4-5: Comparison of seasonal snow depth measurements from the automated SR50a sensor at Strawberry Creek

(dashed line) and the manual measurements from the Waterloo-Wellington International Airport (solid line).

0.000.050.100.150.200.250.300.350.400.450.50

21

/11

/20

08

28

/11

/20

08

05

/12

/20

08

12

/12

/20

08

19

/12

/20

08

26

/12

/20

08

02

/01

/20

09

09

/01

/20

09

16

/01

/20

09

23

/01

/20

09

30

/01

/20

09

06

/02

/20

09

Sno

wd

ep

th (

m)

SR50a W-W

Page 74: Multiscale modelling of snow depth over an agricultural

56

4.3.2 Meteorological data pre-processing in order for proper model runs

As with many models, the CRHM platform requires specific formats of input data. Therefore, in

order to correctly run the model some of the meteorological data format had to be pre-processed.

Time and date stamps were combined into a decimal value, solar radiation was split up into

incoming and outgoing short and longwave radiation, the windspeed was adjusted from a 2 m

reference height to a 10 m reference height, solid precipitation was averaged on a daily basis, and

the vapour pressure was derived within the model using an observation filter from values of

temperature and relative humidity. Data from the SR50a sonic ranger was only used if it had a

signal quality between the range of 152 to 210. Signal quality of 151 and below resulted in the

sensor‟s inability to read distance. Signal qualities between 211 and 300 and 300+ had reduced

echo signal strength and high measurement uncertainties respectively, and were therefore

discarded. Any missing data was replaced with the preceding data of good quality.

4.4 Snow Surveys

A systematic linear transect sampling strategy was adopted to sample the mean, variance, and

distribution of snow throughout the entire field site. Snow depth surveys were conducted over

the winter of 2008-2009 using a GPS-magnaprobe from Snowhydro®. Surveys were conducted

on a winter storm basis or a bi-weekly basis if no new accumulation of snow had fallen since the

previous storm. Surveys were completed on the days of Dec. 08, in 2008 and Jan. 15, 21, 29,

Feb. 04, in 2009. Gaps in the surveying structure were due to the fact that several melt periods

occurred where there was no longer any snow on the ground surface to be measured. This

occurred twice over the entire season, once in the middle of December and once again in the

middle of February.

Page 75: Multiscale modelling of snow depth over an agricultural

57

Initially, only two transects were chosen to cover the entire field site; a NW-SE transect and a

NE-SW transect creating an „X‟ pattern across the site. After much review, it was decided to

include a perimeter transect as well as two more parallel transects (N-S) in order to better

represent snow depth through spatial interpolation. Perimeter transects were labelled in order of

geographical location (ie. Northern, Eastern, Southern, Western, Northwestern) and the interior

transects were labelled numerically (ie. 1, 2, 3, 4). The following diagram (Fig. 4-6)

demonstrates the pattern of snow depth transects used. Snow depths were measured at 5-m

intervals along these transects.

Figure 4-6: Map showing field site and snow depth transects

When using the Magnaprobe, the depth accuracy is ± 3mm if accurately calibrated but this is

conditional on snow conditions, underlying vegetation, and battery power. On several occasions

Page 76: Multiscale modelling of snow depth over an agricultural

58

when there were only trace amounts of snow on the ground, no surveys were conducted because

the Magnaprobe was unable to detect these small amounts. The accuracy may be ±3mm but that

is only when there is enough snow at the right density threshold to allow the depth sensor to

move along the rod and record a value. The GPS accuracy (both absolute and relative) has been

shown to be ± 5m but is dependent on atmospheric conditions and satellites availability. This did

not prove to be an issue at this field site. Data from the MagnaProbe is stored on a Campbell

Scientific CR800 datalogger and is downloaded for analysis using Campbell Scientific‟s

LoggerNet software package.

4.5 Model Experiments

In addition to field sampling, PBSM and MSM modules were used within the CRHM model to

estimate the snow accumulation and distribution over the entire field site. The model was driven

with half-hourly averaged air temperature, wind speed, humidity, incoming and outgoing short

and longwave radiation, and daily averaged solid precipitation data during the snow

accumulation period, 21 November 2008 to 12 February 2009, (as was derived from the SR50a

data). It is assumed that all forcing data were uniform across the entire field site. The model runs

were divided into two different methods, lumped and distributed, respectively. Furthermore, the

lumped method experiments were sequentially run at two different scales; point scale and field

scale. The distributed method was run solely at the field scale.

4.5.1 Lumped Model – Point Scale

Initially, the model was run in the lumped format at the point scale using the forcing data from

the meteorological station and then compared to an average of 10 observed snow depth points

Page 77: Multiscale modelling of snow depth over an agricultural

59

from the same location and timestamp. Model runs at this scale assume a single HRU

(approximately 0.025 km2).

4.5.2 Lumped Model – Field Scale

Secondly, the model was scaled up in lumped format to the field scale using the same forcing

data and then compared to observed snow depth points from the entire field site for each date of

survey. Model runs at this scale also assume a single HRU (approximately 0.406 km2).

4.5.3 Distributed Model – Field Scale

In order to test the model‟s ability to spatially discretize, the model was then run at the field scale

but in a distributed format. In order to do so, the field site was subdivided into three separate

HRUs named Field 1 (0.183km2), Stream (0.011km

2), and Field 2 (0.212km

2) as seen in Figure

4-7 and Table 4-2, and modelled results are compared to observed snow depth points gathered

from each HRU.

4.5.4 Model Parameters

In order to adapt the model for the Strawberry Creek field site, modifications were made to the

model parameters. Control case parameters include the basin area being set to 0.4054 km2, the

basin and observation elevation being set to 343 m asl, latitude set for 43.5°N, observation

reference height (Zref) set to 2.47m above the ground surface, and the vegetation height set to

0.05m. Due to the rolling terrain of the field site, the blowing snow fetch distance was assumed

to be the PBSM minimum of 300 m. All other parameters were set to default (Table 4-2,

Appendix 1). Based on the literature (Pomeroy et al, 1993) and knowledge of the terrain, the

model snow density was assumed to be 250 kg m-3

. Distributed method parameters include

Page 78: Multiscale modelling of snow depth over an agricultural

60

adjusting the HRU areas to those specified above and vegetation heights of 0.1m, 1.5m, and

0.05m respectively for Field 1, Stream, and Field 2. These values were determined based on in

situ observations. The value of 1.5m of vegetation height for the stream was determined to best

represent the combination of the vegetation height in the riparian zone as well as the topographic

depression of the stream itself.

Figure 4-7: Distributed map of CRHM hydrological response units Field 1 (green), Stream (blue), and Field 2 (pink)

Table 4-2 demonstrates all of the model parameters and their values. A definition of each of

these parameters can be found in Appendix 1. Sheltering effects of the farm buildings, driveway,

and garden vegetation were not considered in any of the model runs.

Page 79: Multiscale modelling of snow depth over an agricultural

61

CRHM Module

Parameter Lumped Distributed

HRU1 HRU2 HRU3

Bas

in

basin_area 0.4054

basin_name SC

hru_area 0.4054 0.1827 0.01062 0.2121

hru_ASL 0 0 0 0

hru_elev 343 343 343 343

hru_GSL 0 0 0 0

hru_lat 43.5 43.5 43.5 43.5

hru_names HRU Field 1 Stream Field 2

Run_ID 1

OB

S

catchadjust 0 0 0 0

HRU_OBS 1 1 1 1

lapse_rate 0.75 0.75 0.75 0.75

Obs_elev 343 343 343 343

ppt_daily_distrib 1 1 1 1

tmax_allrain 0 0 0 0

tmax_allsnow 0 0 0 0

MSM

_Org

a1 1.08E+07 1.08E+07 1.08E+07 1.08E+07

a2 7.20E+05 7.20E+05 7.20E+05 7.20E+05

amax 0.84 0.84 0.84 0.84

amin 0.5 0.5 0.5 0.5

smin 10 10 10 10

Z0 snow 0.01 0.01 0.01 0.01

Zref 2.47 2.47 2.47 2.47

PB

SM

A_S 0.001 0.001 0.001 0.001

distrib 0 0 1 1

fetch 300 300 300 300

Ht 0.05 0.1 1.5 0.05

inhibit_bs 0 0 0 0

inhibit_evap 0 0 0 0

N_S 320 320 320 320 Table 4-2: CRHM model parameters

Values for the blowing snow distribution factor, Dp, were determined by trial-and-error runs with

the knowledge of geographical proximity and aerodynamic roughness for each HRU.

Page 80: Multiscale modelling of snow depth over an agricultural

62

HRU name Area (km2) Vegetation Height

(m)

Blowing snow

fetch distance (m)

Blowing snow

distribution factor

Dp

Field 1 0.183 0.1 300 2

Stream 0.011 1.5 300 10

Field 2 0.212 0.05 300 1

Total field site 0.405 0.05 300 1 Table 4-3: Characteristics of HRU parameters for blowing snow experiments at Strawberry Creek

In Table 4-3, a Dp value of 1 and 2 was assigned to Field 2 and Field 1 HRUs respectively

because they both have low aerodynamic roughness and are typically blowing snow „sources‟. A

Dp value of 10 was assigned to the Stream HRU due to it being a topographic depression with a

high aerodynamic roughness and therefore acting as a blowing snow „sink‟. For the lumped

model experiment, a Dp value of 1 was given for the entire field site.

A schematic diagram depicting the flow of inputs and outputs between the modules within

CRHM can be seen in Fig. 4-8. In addition to the MSM and PBSM modules, basin (BASIN),

observation (OBS) and interception (INTCP) modules were added as they were required for

complete model construction. The BASIN module allows the model user to set the parameters as

required for each HRU being used. Meteorological forcing data is put into the OBS and MSM

modules where it is filtered into specific HRU observations and used for the generation of

snowmelt energy values respectively. The INTCP module is required by PBSM to deal with any

incoming precipitation being held within a canopy. In the case of this thesis, there is no canopy

to deal with therefore, INTCP is negligible and merely acts to change the name of incoming

Page 81: Multiscale modelling of snow depth over an agricultural

63

t

rh

ea

u

p

ppt

QSi

Qli

hru_t

hru_rh

hru_u

hru_p

hru_rain

hru_snow

hru_newsnow

net_snow

SWE

t

u

rh

*P

Snowmelt

energy

Figure 4-8: A simplified schematic diagram of the progression of module input

observations through the CRHM model modules.

OBS

INTCP

MSM_Org

PBSM Snow depth

BASIN

Page 82: Multiscale modelling of snow depth over an agricultural

64

precipitation between OBS and MSM/PBSM. This is a necessary step in order for the

MSM/PBSM modules to reckognize the precipitation. The *P denotes a backwards PUT variable

which is an input variable that is derived from another module. In this case, the MSM_Org

module requires both SWE and net_snow from PBSM and INTCP before any snowmelt energy

values are generated. Once snowmelt energy values are generated, they are input into PBSM and

a final value for snow depth is achieved.

4.6 Model testing and verification

Initially, model results of snow depth are compared to measured snow depth for each day of

snow surveying and then compared throughout the entire season. This creates a temporal analysis

that enables testing of the model performance on a daily and seasonal basis. Spatial analysis is

accomplished by analyzing the model results between the various scales of modelling. According

to Rykiel (1996), “validation is a demonstration that a model within its domain of applicability

possesses a satisfactory range of accuracy consistent with the intended application of the model.”

In order to uphold this criterion, the purpose, performance criteria and context of the model are

all well defined. The purpose of the model is to simulate the seasonal snow accumulation

occurring at Strawberry Creek during the winter of 2008-2009. The performance criteria include

the production of a quantitative agreement between observed and simulated snow depths. The

model context was limited to the Strawberry Creek field site and the winter of 2008-2009 for

three specific reasons: 1) the vegetation and topography represent a prairie-like condition on a

small scale in as much as the field was not sheltered significantly by trees or relief, 2) the snow

accumulation and melt periods are well defined and easily determined, and 3) the forcing

meteorological data is year specific. Qualitative comparisons were made by visual inspection of

Page 83: Multiscale modelling of snow depth over an agricultural

65

modelled and observed results of snow depth in graphical formats. The evaluation of the model

performance is done by three statistical approaches, Root Mean Square Difference (RMSD),

Nash-Sutcliffe coefficient (NS) (Nash and Sutcliffe, 1970) and Model Bias (MB), which are

calculated as:

[4.2]

[4.3]

[4.4]

Where n is the number of samples, SDo, SDs, and are the observed, simulated, and mean of

the observed snow depths, respectively. The RMSD is a weighted measure of the difference

between simulated and observed snow depths and has the same units of measurement. The NS

coefficient measures the model efficiency and MB supports NS by quantifying the model‟s under

and overprediction.

As the number of environmental models have increased over the past few decades, so too has the

need for the development of more precise and accurate estimates of the desired variable(s) from

the model results. This in turn led to an increase in the use of a variety of different statistical

methods assessing accuracy or goodness of fit and error (as seen in recent papers: Fekete et al.,

2004; Cavazos and Hewiston, 2005; Willmott and Matsuura, 2005). Some of the more popular

and widely reported error measures found in the literature include coefficients of correlation

(also known as Pearson‟s product moment correlation; r) and determination (r2) as well as the

root mean square deviation (RMSD). There is much debate over which statistical method

provides the best evaluation of model accuracy and selecting a method can be difficult seeing as

Page 84: Multiscale modelling of snow depth over an agricultural

66

how it is based not only on the researcher‟s personal preference but also on the type of model

being used, the method in which the model was calibrated, and the amount of estimated data

points being compared against observed data. W.R. Black (1991) argues that the coefficients of

correlation and determination are measures of relationship or association between estimated and

observed data but they are not the best measure of accuracy or goodness of fit when comparing

model estimates with observed data. Instead, RMSD should be used (Black, 1991). Previous

model experiments using PBSM have included RMSD as one of the main methods of assessing

model accuracy/error and it seemed only fitting to use the same method in this thesis. Even

though RMSD is one of the most widely reported error measures in the literature, it is not

without its own inherent error (see Willmott and Matsuura, 2005). RMSD is not standardized,

making it difficult to look at on its own as a sole measure of model performance. Even though it

demonstrates an amount of average error in the same units as the variables being estimated, this

error value is only a relative term. It is for this reason that other statistical methods were selected

(Nash-Sutcliffe coefficient (NS) and the Model Bias (MB)) in addition to RMSD to demonstrate

the model‟s performance in this paper. NS is widely used and reliable statistic for assessing

goodness of fit of hydrological models. It is very similar to the Pearson product in that the values

of the coefficient give a measure of correlation and can be anywhere between - and +1.

However, unlike the Pearson product, which is theoretically only applicable to linear models that

include an intercept and is greatly affected by model bias, NS can be applied to a variety of

models regardless of model bias (Nash and Sutcliffe, 1970; McCuen et al., 1990, 2006). With

NS, a value of 1 implies that the model perfectly predicts the simulated results compared to the

observed. A value equal to zero indicates that the simulated results are no better than the average

of the observed values. A value less than zero indicates that the observed mean is a better

Page 85: Multiscale modelling of snow depth over an agricultural

67

predictor then the simulated results. Therefore, any positive value of this coefficient indicates

that the model has some predictive power and the higher the value, the better the model

performs. Model Bias (MB) is added to support the predictive power of the model as seen in the

Nash-Sutcliffe coefficient. A positive value and negative value of MB indicate overprediction

and underprediction of the model, respectively. Following similar research from Fang and

Pomeroy (2009), an acceptable range of model performance was chosen to be <30% (MB <

0.30).

4.7 Summary

This chapter provides a description of the methodology and modelling approach used in this

thesis. It provides a detailed description and geological history of the study site as well as all of

the data used to map out the study site and snow surveys. All methods of acquiring

meteorological data and the required alterations to this data used in the model experiments are

discussed in length. Snow surveys play a key role in validating the model experiments and this

section describes in full the steps taken to acquire snow depth values across the entire field site

using a GPS-magnaprobe throughout the winter season of 2008-2009. Model experiments as

well as model validation and testing processes are also outlined. The final section justifies the

statistical methods used to assess the accuracy or goodness of fit of the model compared to the

observed data.

Page 86: Multiscale modelling of snow depth over an agricultural

68

Chapter 5.

Results

5.1 Field Results

The precipitation pattern at Strawberry Creek during the winter season of 2008-2009 featured a

dry fall and early winter with snow accumulating on the ground in late November. Most of the

snow accumulation in this area occurs between the months of December and March. In some

anomalous years snow accumulation can begin as early as October and remain as late as April.

Most of the snow accumulation occurs due to several large precipitation events throughout the

entire winter season which was the case for this particular winter season. Table 5-1 demonstrates

monthly averages from Strawberry Creek compared to the monthly climatic average from the

Waterloo-Wellington International Airport located in Breslau, 10km to the southwest.

Variable Strawberry Creek Waterloo-Wellington Airport

Nov. Dec. Jan. Feb. Nov. Dec. Jan. Feb.

Temp (C) -2.96 -4.92 -10.39 -5.17 2.30 -3.80 -7.10 -6.40

Rel. Humidity (%) 97.27 98.93 97.86 96.41 87.50 87.00 86.40 83.60

Wind Speed (m/s) 2.62 4.02 2.96 3.23 3.94 4.06 4.28 3.89

Wind Direction (deg) 195.91 212.04 216.85 210.15 225.00 225.00 225.00 270.00

Snow depth (m) 0.05 0.12 0.26 0.36 0.01 0.05 0.12 0.14

Table 5-1: Calculated monthly climatic averages for Strawberry Creek during the 2008-2009 winter season compared to

1971-2000 climatic averages from the Waterloo-Wellington Airport (Environment Canada, 2010)

Climatic averages for temperature, relative humidity, and snow depth were all above the long

term average for the area. Wind speed was not as strong and the wind direction came from a

more southerly direction then the long term average. Collected meteorological data from

Strawberry Creek can be seen in Fig 5-1.

Field observations demonstrate that the Strawberry Creek field site has a characteristic and

predictable snow distribution pattern associated with wind direction, wind speed, topography,

Page 87: Multiscale modelling of snow depth over an agricultural

69

Date

Figure 5-1: Corrected meteorological data for Strawberry Creek from November 21, 2008 to February 12, 2009

-25.000

-20.000

-15.000

-10.000

-5.000

0.000

5.000

10.000A

ir T

emp

(°C

)

0.000

20.000

40.000

60.000

80.000

100.000

Rel

ativ

e H

um

idit

y (%

)

0.000

2.000

4.000

6.000

8.000

Win

d S

pee

d (

m/s

)

0

100

200

300

Win

d D

irec

tio

n (

deg

.)

0.00

0.02

0.04

0.06

0.08

0.10

20/11/2008 30/11/2008 10/12/2008 20/12/2008 30/12/2008 09/01/2009 19/01/2009 29/01/2009 08/02/2009

Sno

wfa

ll (m

)

Page 88: Multiscale modelling of snow depth over an agricultural

70

and vegetation. During the 2008-2009 winter season, the wind direction typically came from the

south southwest (196 to 217) with an average speed of 3 m s-1

(Table 5-1, Fig. 5-1).

Definitive snow accumulation patterns appeared across the entire field site with drifts occurring

around the edges of vegetation stands and within the creek banks. Scouring occurred in the open

tilled farmlands (N and NE fields) which have a low snow-holding capacity. The SW field had a

slightly higher snow-holding capacity due to the presence of the winter grass crop and therefore

had an increase in snow depth compared to the other two farm fields. The forested borders along

the eastern and western edges of the field site reduced the wind speed substantially, effectively

acting as snow fences, facilitating the accumulation of large quantities of snow (40 to 90 cm

deep). Within the creek banks the snow drifts would completely fill the banks (100 cm) in some

areas and at times, even pile higher, pushing the magnaprobe to its depth limit of 120 cm.

Transect profiles demonstrating snow depth and elevation can be seen in Appendix 2. Fig. 5-2 is

a graphical representation of both the observed snow depth averages on the days of snow

surveying and the modelled snow depth results. It shows the gradual increase in mean field-scale

snow depth from transect data throughout the winter season.

The extreme values in the data are due to snow depth readings from within the creek banks and

drifts along the field site borders. The black circles represent the mean snow depth and the upper

and lower box limits represent the 1st and 3

rd quartiles respectively.

5.2 Model Results

Two modelling approaches were taken in this study: Lumped and distributed. The result of these

approaches can be seen in Fig 5-2 and 5-3 respectively. In both approaches, the model failed to

simulate the early onset of snow until the middle of December and then steadily simulated the

Page 89: Multiscale modelling of snow depth over an agricultural

71

seasonal snow accumulation throughout the rest of the test season, which as was previously

mentioned, fell between November 21, 2008 and February 12, 2009. The lumped approach

reached a maximum snow depth of 33.0 cm on two different occasions in the month of January.

In the distributed approach the maximum snow depth from the HRUs Field 1, Field 2, and

Stream were 29.0cm, 28.0cm, and 41.0 cm respectively and occurred at the middle and end of

January.

Figure 5-2 Comparison of averaged model (lumped) and observed snow depth during the 2008-2009 winter season. The

snowfall graph is added to demonstrate snowfall activity throughout the season.

0.00

0.02

0.04

0.06

0.08

0.10

Sno

wfa

ll (m

)

0

0.2

0.4

0.6

0.8

1

1.2

21

/11

/20

08

28

/11

/20

08

05

/12

/20

08

12

/12

/20

08

19

/12

/20

08

26

/12

/20

08

02

/01

/20

09

09

/01

/20

09

16

/01

/20

09

23

/01

/20

09

30

/01

/20

09

06

/02

/20

09

Sno

wd

ep

th (

m)

Modelled SD(m) Observed SD(m)

Page 90: Multiscale modelling of snow depth over an agricultural

72

Figure 5-3 Comparison of averaged simulated model snow depth and cumulative snowfall (lines) and observed snow

depth (symbols) for each HRU in the distributed approach over the 2008-2009 winter season.

Accumulation of snow depth in all three HRUs fall predominantly below the cumulative

snowfall line which suggest that the entire field site is subjected to the effects of erosion from the

wind and therefore acts as a source of blowing snow. The only exception to this occurs in the

stream HRU between the end of December and the middle of January. In this case, the stream

acts as a blowing snow sink.

5.3 Comparison of Simulated Model Results and Observed Snow Survey Results at the Point

and Field Scales.

The simulated model results and field observations of snow depth were compared to one another

in three different methods. The first two methods make use of model results from the lumped

modelling approach and compare them to field observations at the point and field scales

respectively. The third method makes use of model results from the distributed modelling

0.0

0.1

0.2

0.3

0.4

0.5

0.6

21

/11

/20

08

28

/11

/20

08

05

/12

/20

08

12

/12

/20

08

19

/12

/20

08

26

/12

/20

08

02

/01

/20

09

09

/01

/20

09

16

/01

/20

09

23

/01

/20

09

30

/01

/20

09

06

/02

/20

09

Sno

wd

ep

th (

m)

Field 1 Stream Field 2 Field 1 Stream 2 Field 2 Snowfall

Page 91: Multiscale modelling of snow depth over an agricultural

73

approach and compares them with field observations which have been collected from each of the

three HRUs used in the model (Field 1, Stream, and Field 2). Table 5-2, below, displays the

sample size (n value) used in the statistical analysis for the observed snow depth for each day of

survey.

Date

Sample size (n)

Lumped Distributed

Point Scale Field Scale Field 1 Stream Field 2

Dec. 08/08 20 1138 426 44 668

Jan. 15/09 20 856 340 40 476 Jan. 21/09 20 1687 874 112 832 Jan. 29/09 20 1346 614 90 642

Feb. 04/09 20 1280 638 56 638 Table 5-2 Sample sizes of snow depth for each day of snow survey used for statistical analysis

5.3.1 Point Scale Model Results: Lumped

To quantify the performance of the model‟s ability at predicting snow depth at the point scale,

RMSD, NS, and MB we computed for each individual date of snow survey during the 2008-2009

Date RMSD (m) NS MB

Dec.08/08 0.074 -0.097 -0.943 Jan. 15/09 0.051 0.650 0.532 Jan. 21/09 0.022 0.969 0.160 Jan. 29/09 0.004 0.999 -0.027

Feb. 04/09 0.024 0.978 -0.135 Table 5-3 Root Mean Squared Deviation (RMSD), Nash-Sutcliffe coefficient (NS), and Model Bias (MB) for the

comparison of point-scale simulated snow depth in a lumped approach and observed snow depth for each date of snow

transect surveys during the 2008-2009 winter season.

winter season (Table 5-3). RMSD and NS values range from 0.004 m to 0.074 m and -0.097 to

0.999 respectively. The model performs well on Jan. 21/09 and Feb. 04/09 and performs really

well on Jan. 29/09. As was previously mentioned, the model performed poorly in the early half

of the season which is evident in the statistical analysis, most notably on Dec. 08/08 where the

RMSD value was 0.074 m and the NS value was -0.097. The amount of overestimation and

Page 92: Multiscale modelling of snow depth over an agricultural

74

underestimation for total snow depth is within 16% for the latter three survey days which is well

within an acceptable range of 30% (Fang and Pomeroy, 2009), however, the model

underestimates by as much as 94% on the first survey day and then overestimates by more than

52% on the following survey day which is unacceptable.

5.3.2 Field Scale Model Results: Lumped

Upscaling from the point scale the field scale is accomplished by including observations from the

entire field site instead of only those which are in close proximity to the meteorological station.

Date RMSD (m) NS MB

Dec.08/08 0.047 0.163 -0.913 Jan. 15/09 0.056 0.600 0.631 Jan. 21/09 0.014 0.977 0.150 Jan. 29/09 0.010 0.995 -0.068

Feb. 04/09 0.023 0.978 -0.147 Table 5-4 Root Mean Squared Deviation (RMSD), Nash-Sutcliffe coefficient (NS), and Model Bias (MB) for the

comparison of field-scale simulated snow depth in a lumped approach and observed snow depth for each date of snow

transect surveys during the 2008-2009 winter season.

By incorporating more field observations, there is an averaging effect on the statistical analysis

(Table 5-4). Once again the model performs well on the final three days of snow surveys, most

notably on Jan. 29/09, with RMSD values for those three days ranging from 0.010 m to 0.023 m

and NS values ranging from 0.977 to 0.995. The model still performs poorly on the first two

survey days but performs slightly better at the field scale then at the point scale. The amount of

overestimation and underestimation is within 15% at the field scale which is slightly closer to the

observations than the point scale modeling results over the last three days of snow surveys. The

model underestimates snow depth by 91% on the first survey day which is slightly better than the

point scale results but then over estimates by 63% on the second survey day which is worse than

the result at the point scale. The faults of the results are likely due to uncertainty in the modelling

of blowing snow over both complex terrain and varying vegetation.

Page 93: Multiscale modelling of snow depth over an agricultural

75

5.3.3 Field Scale Model Results: Distributed

The final method compares field observations that were assigned to each of the HRUs with

simulated snow depths using a distributed modelling approach at the field scale for each of the

five days of snow survey. RMSD, NS, and MB were calculated for each HRU and can be seen

below in Table 5-5.

Date RMSD (m) NS MB

Field 1 Stream Field 2 Field 1 Stream Field 2 Field 1 Stream Field 2

Dec. 08/08 0.058 0.162 0.033 0.132 -0.039 0.218 -0.928 -0.973 -0.881 Jan. 15/09 0.036 0.027 0.042 0.848 0.968 0.736 0.387 0.170 0.511 Jan. 21/09 0.012 0.020 0.010 0.992 0.987 0.994 0.088 0.111 0.078 Jan. 29/09 0.032 0.011 0.022 0.965 0.997 0.981 -0.186 -0.051 -0.139 Feb. 04/09 0.048 0.067 0.040 0.926 0.929 0.941 -0.271 -0.258 -0.242

Table 5-5 Root Mean Squared Deviation (RMSD), Nash-Sutcliffe coefficient (NS), and Model Bias (MB) for the

comparison of field-scale simulated snow depth in a distributed approach and observed snow depth for each HRU on

each date of snow transect surveys during the 2008-2009 winter season.

There is a large variation in the values of RMSD and NS which range from 0.010 m to 0.162 m

and -0.039 to 0.997 respectively. Although the model still fails to adequately simulate snow

depth during the initial snow accumulation period, the distributed approach does a better job than

the lumped approach on the second day of snow surveying and continues to perform well

throughout the remainder of the season. By examining the model performance for each HRU on

each survey date we can see that the model does not perform well at all on Dec. 08/08. On Jan.

15/09 the model still has trouble simulating snow depth in both Field 1 and Field 2 but does well

simulating it in the Stream, only overestimating by 17%. By Jan. 21/09 the model is performing

very well in all three HRUs and is overestimating by less than 11%. The model begins to lose

performance by Jan. 29/09 but is still performing well, especially in the Stream. The model is

now underestimating snow depth by less than 19% but is still within the acceptable range as

mentioned above. By the final survey date, Feb. 04/09, the model continues to lose performance

Page 94: Multiscale modelling of snow depth over an agricultural

76

and underestimates snow depth in all three HRUs. This underestimation ranges from 24-27%

which although is worse than the previous survey date, is still within the acceptable range of

model performance. On the whole, the model performs better for the two field HRUs during the

latter three days of surveying but performs better during the middle three days for the Stream

HRU. Examinations of MB values indicate that the model is able to generate mid to late season

snow depth values that are moderately close to the observations, keeping well within the

acceptable range of 30%.

5.4 Summary

Field and model results are examined in this chapter. Field results of both meteorological and in

situ snow surveys demonstrate that the 2008-2009 winter season was above average for

temperature, relative humidity, and snow depth but was below average for wind speed, resulting

in a shorter than average winter season with larger amounts of snow depth and less than average

redistribution of snow. A constant wind direction allows for a definitive and predictable pattern

of snow distribution across the entire field site. Examination of cumulative snowfall and snow

depth suggest that the field site acts predominantly as a blowing snow source throughout the

majority of the winter season with the exception of the stream HRU acting as a blowing snow

sink from the end of December to the middle of January.

The model appears to perform fairly well in both the lumped and distributed approaches once a

stable base layer of snow is established on the ground. Early seasonal snow accumulation is not

modelled well no matter what approach is taken at any scale. The model also begins to lose

performance at the tail end of the season, specifically in the distributed approach but is still

performing at an acceptable range. When comparing the point scale with the field scale in the

Page 95: Multiscale modelling of snow depth over an agricultural

77

lumped approach, the model performs almost identically with only a few minor differences.

When comparing the lumped approach to the distributed approach the performance is similar but

is spatially more accurate in the distributed approach due to the spatial discretization of this

modelling approach. Reasons for the poor model performance and sources of error will be

discussed further on in the discussion section of this thesis.

Page 96: Multiscale modelling of snow depth over an agricultural

78

Chapter 6.

Discussion and Conclusions

6.1 Evaluating Model Performance and Sources of Error

Even though the model represented the average snow depth fairly well for most of the winter

season, there are many sources of error that influenced the outcome of the model which

prevented a better outcome. The following section discusses the model performance and sources

of error that affected the outcome of the model results. Some of the topics include proper terrain

representation, poor early-season simulation, model limitations, and adjustments to

meteorological forcing data.

Terrain representation was a challenge with this model. The model platform used in this study

assumed a flat terrain and did not allow for the incorporation of a digital elevation map or any

changes to topography within the basin. The only exception was the ability to adjust the HRU

elevation. Upon initial testing however, this proved to be effective only when the change in

elevation was significant (>15m) and over a significantly larger area than the HRUs created in

this study. Therefore it was assumed that the field site was flat. In actuality however, the total

change in elevation was 9 m and the field site had several topographic features including hills,

the creek, and a sloping terrain on either side of the creek. In addition to the natural topography

of the field site, there were many objects (both man-made and natural) that added to the

difficulty in accurately modelling the terrain. These objects included the many farm buildings

located at the centre of the field, the elevated driveway, and all of the trees that bordered the

property, driveway, and buildings. All of these variations in terrain create changes in the

distribution of snow as it accumulates on the ground throughout the winter season. The hills are

important areas for erosion and deposition of blowing snow on both the windward and leeward

Page 97: Multiscale modelling of snow depth over an agricultural

79

sides respectively. The sloped terrain helps to channel saltating and suspended snow to areas of

lower topography, namely the creekbed which in turn, was the largest sink of blowing snow

across the entire field site. The steep banks of the narrow creek allowed for deep drifts to

accumulate on either side and in some instances, completely fill the 1.5 m deep creekbed,

creating little snow bridges across several sections of the creek. The farm buildings provide a

significant amount of sheltering effect and a fair amount of drifting was observed in the lee of

these buildings. However, due to the proximity of the buildings to the creek, much of the drifting

was contained within the creek bed and did not progress significantly into the downwind field.

The elevated driveway produces a small amount of drifting on either side. The tall trees that

surround the buildings and run along portions of the elevated driveway are considered to be

within the sheltering effects of these man-made objects and therefore did not significantly alter

the drifting that occurred. The trees that are located along the perimeter of the property act as

wind barriers and allow for the accumulation of deep snow drifts around their bases. Air flow

through these trees is complex and difficult to model realistically. Previous studies have been

conducted to observe the relationship between trees and snow including the effects of shelterbelt

trees and shrubs (Sturges, 1983) and snow fences (Tabler, 1980a). The version of CRHM used in

this study does have a module that can incorporate the effect of canopy interception and wind

flow through trees but it was not incorporated because it involves a much greater amount of input

variables which was far outside the realm of this study. Furthermore, without the coupling of

this model to a 3-D windflow and terrain model such as the Mason and Sykes three-dimensional

extension of the Jackson and Hunt theory (MS3DJH) terrain wind-flow model (Walmsley et al.,

1986), correct terrain model representation is unobtainable, therefore leading to increased error.

Page 98: Multiscale modelling of snow depth over an agricultural

80

Secondary to the topic of proper terrain modelling would be the inclusion of an updated working

GIS interface. The version used in this thesis was ArcGIS/ArcView 9.3 but the model was

configured to use ArcView version 3.1 and would not allow for more than one HRU shape file to

be uploaded at a time. Personal conversation with the model developer, Tom Brown, from the

University of Saskatchewan has determined that this is an area that needs to be updated and

improved upon.

The model does not simulate the early onset of snow and this can be attributed to a combination

of module assumptions from both PBSM and MSM and the inability to accurately model

physical processes at a fine scale in the natural environment. As was mentioned above, the early

winter season had several small snowfall and melt periods which are very difficult to accurately

model. The MSM module negates any effect of energy coming from the ground, from

precipitation, and internally from within the snow pack, thus altering the energy and mass

balance equations. In the early winter season the heat energy coming from the substrate and from

rain can significantly affect the overlying snowpack. Internal changes to the snowpack are

minimal due to the presence of such a small amount of snow on the ground unless significant

snowfall has occurred in various phases, resulting in a stratified snowpack. Nevertheless, MSM

is unable to simulate these small periods of melt during the early part of the winter season.

PBSM sequentially distributes any transported snow from areas of low snow holding capacity to

areas of higher snow holding capacity regardless of wind direction. Although this method works

in theory, it is not always the case in the natural environment. As well, snow transport within

PBSM only occurs once the first HRU has reached its snow holding capacity. This becomes an

issue in the beginning of the winter season in both the lumped and distributed approaches. In the

lumped approach, a snow depth amount of 5 cm is needed before any snow transportation can

Page 99: Multiscale modelling of snow depth over an agricultural

81

occur. In the distributed approach an amount of 5 cm and 10 cm is necessary to accumulate in

Field 2 and Field 1 respectively before any snow is distributed to the Stream. In addition to the

minimum snow depth requirements, the sequence in which snow transport occurs between HRUs

does not represent the natural process very well. Since there is no method of accounting for wind

direction within the module, PBSM distributed snow based solely on snow-holding capacity. In

the case for Strawberry Creek snow distribution would follow in the format of Field 2 to Field 1

and then into the Stream. This is not the natural format of the field site (as seen in Figure 4-6).

As well, the prevailing wind direction actually comes from the west, resulting in a natural

sequence of snow distribution from Field 1 to Stream and then to Field 2. In addition to the

above mentioned sources of error, another significant factor is the inability to simulate accurate

amounts of snow depth in the initial HRU due to the fact that data from outside of the study

domain cannot be captured before it enters the study domain.

Once however, all (or the majority) of the snow-holding capacities have been reached, the model

performs really well at simulating the average snow depth across the entire field site, as seen in

the model results section. The snowpack becomes one homogenous layer and snow transport is

allowed to flow naturally between the HRUs. This is evident in the middle of the winter season.

The end of season snowmelt underestimation is primarily due to the previously mentioned

assumptions found within the MSM module. The HRU that is best represented throughout the

entire season is Field 2. This is due to the fact that is has the lowest snow-holding capacity (5cm)

and any blowing snow is placed into this HRU before moving onto the next. Although the

natural progression would see blowing snow come from Field 1 and the Stream HRUs, or from

outside the study domain, the model‟s ability to represent this HRU well is partially coincidental.

Page 100: Multiscale modelling of snow depth over an agricultural

82

Accounting for precipitation events throughout the winter season can be a challenging

endeavour. Initially, precipitation in the form of rain would be collected using a tipping-bucket

rain gauge and snowfall would be a product of the difference in changes of accumulated snow

depth from the SR50a snow depth sensor. However, after consulting with the model developer,

we came to realize that the current version of the CRHM platform did not allow for separate

precipitation inputs from rain and snow at the same time. It is assumed that the model user would

be using a Geonor precipitation sensor and that the difference between rain and snow would be

determined within the model by applying a set temperature to which precipitation would either

be rain or snow. These parameters are known as „tmax_allrain and tmax_allsnow‟. Unfortunately

this realization did not arise until halfway through the winter season and resulted in all

precipitation data in the form of rain being neglected. This was unfortunate because the heat

energy from rain can be a significant source of melting, especially in Southern Ontario.

The adjustments made to account for snowfall precipitation were necessary as there was no other

accessible method to account for this form of precipitation throughout the winter season. The

daily snow depth values from the SR50a sensor overestimate a bit more than the values from the

Waterloo-Wellington airport but were chosen because they represent values from the actual field

site rather than from a location 10km away. The calculations necessary to acquire the

precipitation data (difference in accumulated snowfall) are simple and straight forward but they

are heavily reliant on the quality of data acquired by the snow depth sensor. Fortunately though,

the sensor had good signal quality (between 152 and 210) 89.9% of the time.

Page 101: Multiscale modelling of snow depth over an agricultural

83

6.2 Future Considerations

If given the task of performing this study over again, a few changes would be made. In terms of

modelling, PBSM would be coupled with the Energy Balance Snowmelt Model (EBSM) (Gray

and Landine, 1987) as well as with a terrain wind-flow model such as MS3DJH (Walmsley et al.,

1986). These models have proven to be very effective in prairie-like conditions.

In terms of instrumentation, a Geonor precipitation gauge and a particle detector would be added

to the list of instrumentation and would significantly increase model simulation results because

snowfall rates, blowing snow events, and proper designation and inclusion of both forms of

precipitation can all be accounted for. As well, CRHM should be updated so that it could

accommodate the newest version of a GIS interface. This would allow for an enhanced

representation of terrain through a more user-friendly interface by simultaneously incorporating

multiple shape files for each HRU as well as digital terrain models. In doing so, the sheltering

effects of farm buildings and surrounding trees could be incorporated and a better representation

of blowing snow redistribution would be possible.

In terms of methodology, the sampling strategy would be kept the same but the inclusion of

several snowpits along each transect where values for SWE and density would be collected.

These values would help in the testing of the model.

Temporal analysis of the model performance indicates that the model does very well at

simulating the average snow depth across the entire season, regardless of the difficulty of

attaining reasonable simulations for the initial onset of snow. This suggests that if given the

correct seasonal data, the model may be used as a tool to compare seasonal averages for any

number of studies across a number of years.

Page 102: Multiscale modelling of snow depth over an agricultural

84

6.3 Conclusions

This thesis was an attempt at modelling snow depth across an agricultural field in a small

subcatchment in Southern Ontario during the 2008-2009 winter season. It incorporated the

modules of the Prairie Blowing Snow Model (PBSM) and the Minimal Snowmelt Model (MSM)

within the Cold Regions Hydrological Model platform. Two different modelling approaches

(Lumped and Distributed) were compared at various scales (point scale and field scale) to a

series of in situ snow depth measurements, collected from snow surveys, which were completed

throughout the winter season using a semi-automated snow depth sensor. The model performs

well to simulate mid season snow depth averages in both the lumped and distributed approaches

but has a hard time dealing with the onset of snow as well as the smaller scale effects of

vegetation and terrain. There was very little change to model performance when upscaled from

the point scale to the field scale. From an experimental point of view, this demonstrates that there

are no scaling effects to be found at this size of domain.

This model is much better suited for large-scaled applications where these small-scale effects are

negligible. When available, it is also best to couple the model with a 3-D terrain model which

will increase the model‟s ability to accurately simulate the conditions across the entire study

domain (Fang and Pomeroy, 2009; Essery and Pomeroy, 2004). Modelling the natural

environment is extremely challenging and unless you are able to factor in all of the possible

variables as they change throughout the entire season at every possible scale there is no way of

reproducing 100% of the natural environment. The solution is to focus on the objective and

attempt to minimize the amount of error being produced. Design the experiment to suit the study

domain. The results posted in this thesis however, demonstrate that it is still possible to simulate

seasonal snow depth at a variety of scales with relative accuracy. The Prairie Blowing Snow

Page 103: Multiscale modelling of snow depth over an agricultural

85

Model was designed for much larger study domains but this thesis proves that with a little

modification to the model and forcing data and if coupled with a snowmelt module, acceptable

results can be generated at a much smaller scale.

Page 104: Multiscale modelling of snow depth over an agricultural

86

References

Bagnold, R. A. (1941). The Physics of Blown Sand and Desert Dunes. Methuen & Co., London.

265 pp.

Bartelt, P., Lehning, M. (2002) “A physical SNOWPACK model for the Swiss avalanche

warning Part 1: numerical model” Cold Regions Science and Technology, 35. No. 3: 123-145

Benson, C.S (1982). “Reassessment of winter precipitation on Alaska’s Arctic Slope and measurements on the flux of windblown snow”. Geophysical Institute, University of Alaska Report UAG R-288, September 1982. Benson, C.S. and Sturm, M. (1993). Structure and wind transport of seasonal snow on the Arctic Slope of Alaska. Ann. Glaciol., 18, 261–267. Bintanja, R. (1998a). The contribution of snowdrift sublimation to the surface mass balance of Antarctica. Ann. Glaciol., 26, 167-173

Brun, E. et al, (1989) "An energy and mass model of snow cover suitable for operational

avalanche forecasting", J. of Glaciol., 35, No. 121: 333-342

Cavazos, T, Hewitson, B.C. (2005) Performance of NCEP-NCAR reanalysis variables in

statistical downscaling of daily precipitation. Climate Research 28:95–107

Chapman. L.J., Putnam. D.F. ( 1984). The physiography of southern Ontario. 270p.

Chow, V.T., Maidment, D.R., Mays, L.W., (1988). Applied Hydrology. McGraw-Hill, New York

Colbeck, S. (1978) Adv. Hydrosci .11. pp 165 – 206.

Déry, S.J. and Taylor, P.A. (2006). Some aspects of the interaction of blowing snow with the

atmospheric boundary layer. Hydrol. Process. 10, 1345-1358 Déry, S.J. and Yau, M.K. (2001). Simulation of blowing snow in the Canadian Arctic using a double-moment model. Bound.-Layer Meteorol., 99, 297–316. Déry, S.J. and Yau, M.K. (2002). The large-scale mass balance effects of blowing snow and surface sublimation. J. Geophys. Res., 107(D23), 4679, DOI:10.1029/2001JD001251. DeWalle, D. R., Rango, A. (2008) “Principles of Snow Hydrology” Cambridge University Press,

Cambridge. pp. 21

Diamond, M., Lowry, W.P. (1953) “Correlation of density of new snow with 700mb

temperature”. Research Paper 1, Snow, Ice and Permafrost Research Establishment. US Army

Corps of Engineers; pp 3.

Page 105: Multiscale modelling of snow depth over an agricultural

87

Dingman, S. L. (2002) “Physical Hydrology” 2nd

ed. Prentice-Hall Inc., Upper Saddle River,

New Jersey

Diskin, H.M.,( 1970) Research Approach to Watershed Modelling, Definition of Terms, ARS and

SCS Watershed Modelling Workshop, Tuscon, Ariz., March

Elder, K., Dozier, J., and Michaelsen, J., (1991). Snow accumulation and distribution in an alpine watershed. Water Res Res, 27: 1541-1552

Environment Canada, (2010) “Canadian Climate Normals or Averages 1971-2000” http://www.climate.weatheroffice.gc.ca/climate_normals/index_e.html Environmental Systems Research Institute, Inc., (1998) ArcGIS 9/ArcView version 9.3.

Redlands, California, U.S.A.

Essery R, Etchevers P. (2004). “Parameter sensitivity in simulations of snowmelt”. Journal of

Geophysical Research 109 (D20111): Doi:10D1029/2004JD005036.

Essery R, Yang ZL. (2001). An Overview of Models Participating in the Snow Model

Intercomparison Project (SNOWMIP). SnowMIP Workshop 11 July 2001. 8th Scientific

Assembly of IAMAS. Innsbruck

Essery, R., Bunting, P., Hardy, J., Link, T., Marks, D., Melloh, R., Pomeroy, J., Rowlands, A.

and Rutter, N.J. (2008). “Radiative transfer modelling of a coniferous canopy characterized by

airborne remote sensing.” Journal of Hydrometeorology. Vol 9: Issue 2, pp. 228-241

Fassnacht, S.R. (2000) “Distributed Snowpack Simulation Using Weather Radar with an

Hydrologic – Land Surface Scheme Model”. PhD Thesis. University of Waterloo.

Fang, X. And Pomeroy, J.W. (2009) “Modelling blowing snow redistribution to prairie

wetlands”, Hydrological Processes. 23: 2557-2569 DOI: 10.1002/hyp.7348

Fekete, B.M., Vörösmarty, C.J., Roads, J.O., Willmott, C.J. (2004) Uncertainties in precipitation

and their impacts on runoff estimates. Journal of Climatology 17:294–304

Flato, G.M., Boer, G.J., Lee, W.G., McFarlane, N.A., Ramsden, D., Reader, M.C., and Weaver,

A.J., 2000: The Canadian Centre for Climate Modelling and Analysis Global Coupled Model and

its Climate. Climate Dynamics, 16, pp. 451-467.

Geiger, R. (1965) “The Climate Near the Ground”. In Boundary Layer Climates. T.R. Oke. 2nd

ed. Routledge

Goerre, S. et al. (2007) “Impact of weather and climate on the incidence of acute coronary

syndromes” International Journal of Cardiology, vol 118, issue 1, pp. 36-40

Goodison, B.E., Ferguson, H.L., McKay, G.A. (1981) “Measurement and data analysis”. In

Handbook of Snow –Principles, Processes, Management & Use. D.M. Gray and D.H. Male

(Eds). Pergamon Press: Oxford, England; pp 191 - 274.

Page 106: Multiscale modelling of snow depth over an agricultural

88

Goodison, B. E., B. Wilson, K. Wu, and J. Metcalfe (1984) An inexpensive remote snow-depth

gauge: An assessment. Proc. 52nd Annual Western Snow Conf., Sun Valley, ID, 188–191.

Goodison, B.E., Louie, P.Y.T. and Yang, D. (1998). “WMO solid precipitation measurement intercomparison, final report”. WMO/TD 872, World Meteorological Organisation, Geneva.

Gordon, C. et al. (2000) “The simulation of SST, sea ice extents and ocean heat transports in a

version of the Hadley Centre coupled model without flux adjustments” Climate Dynamics, 16,

No. 2-3: 147-168

Grand River Conservation Authority, (2008)

http://www.grandriver.ca/index/document.cfm?Sec=12&Sub1=55&Sub2=24

Greeley, R. and Iversen, J.D., (1985). “Wind as a Geological Process on Earth, Mars, Venus and

Titan.” Cambridge University, Cambridge, 333 pp. Grody, N. (1996) “Global identification of snowcover using SSM/I measurements” Geoscience

and Remote Sensing, 34. No. 1. pp 237-249

Harris, M. (1999). Nitrate attenuation in a narrow non-forested riparian buffer zone in an

agricultural watershed in southern Ontario. MES Thesis. Department of Geography and

Environmental Studies, Wilfrid Laurier University. 339p

Hedstrom, N.R., Pomeroy, J.W. (1998) “Measurements and modelling of snow interception in

the boreal forest”. Hydrological Processes 12(10-11): pp 1611 – 1625.

Hiemstra, C.A., Liston, G.E. and Reiners, W.A. (2002). Snow redistribution by wind and interactions with vegetation at upper treeline in the Medicine Bow Mountains, Wyoming, USA. Arctic, Antarctic, Alpine Res., 34(3), 262–273. Hinzman, L.D., Goering, D.J. and Kane, D.L. (1998). A distributed thermal model for calculating soil temperature profiles and depth of thaw in permafrost regions. J. Geophysical Res., 103(D22), 28975–28991.

Jordan, R. (1991) “A one-dimensional temperature model for a snow cover: Technical

documentation for SNTHERM.89”, Spec. Rep. 91-16, U.S. Army Corps of Eng., Cold Reg. Res.

and Eng. Lab., Hanover, N.H.

Karrow, P. F. (1968). Pleistocene geology of the Guelph area. Geological Report 61. Ontario

Department of Mines.

Karrow, P.F. (1974). Till stratigraphy in parts of southwestern Ontario. Geological Society of

America Bulletin 85. 761-768.

Karrow , P. F. (1993). Quaternary geology of the Stratford-Conestogo Area. Report 283. Ontario

Geological Survey.

Page 107: Multiscale modelling of snow depth over an agricultural

89

Kind, R. J. (1990). Mechanics of aeolian transport of snow and sand‟, J. Wind Engrg and Ind.

Aerodynam., 36, 855-866. King, J.C., Anderson, P.S. and Mann, G.W. (2001). The seasonal cycle of sublimation at Halley, Antarctica. J. Glaciol., 47, 1–8. Kobayashi, D. (1972) Studies of snow transport in low-level drifting snow. Contrib. Inst. Low Temp. Sci., Ser A 24, 1-58

Kuusisto, E. (1986) “The energy balance of a melting snow cover in different environments.” In

E. M. Morris, ed. Modeling Snowmelt-Induced Processes. Wallingford, UK: International

Association of Scientific Hydrology Press. (IAHS Publication No. 155)

La Chapelle, E. (1961) “Snow Layer Densification” Alta Avalanche Study Center, Project F,

Progress Report No. 1, US Department of Agriculture Forest Service, Wasatch National Forest.

Larson, L.W. and Peck, E.L. (1974). Accuracy of precipitation measurements for hydrologic modeling. Water Res. Res., 10, 857–863. Lee, L. W. (1975). “Sublimation of snow in a turbulent atmosphere”, PhD Thesi.7, University of

Wyoming, Laramie, Wyoming, 162 pp. Liston, G.E. (1999). Interrelationships among snow distribution, snowmelt, and snow cover depletion: implications for atmospheric, hydrologic, and ecologic modeling. J. Appl. Meteorol., 38(10), 1474–1487. Liston, G., M. Strum (1998). “A snow-transport model for complex terrain” Journal of Glaciology, 44. No. 148: 498-516. Liston, G.E. and Sturm, M. (2002). Winter precipitation patterns in arctic Alaska determined from a blowing-snow model and snow-depth observations. J. Hydrometeorol., 3, 646–659. Liston, G.E. and Sturm, M. (2004). The role of winter sublimation in the Arcic moisture budget. Nordic Hydrol.,35 (4-5), 325-334 Liston, G.E., McFadden, J.P., Sturm, M. and Pielke, R.A., Sr. (2002). Modeled changes in arctic tundra snow, energy, and moisture fluxes due to increased shrubs. Global Change Biol., 8, 17–32. Male, D.H. (1980) “The Seasonal Snowcover”. In Dynamics of Snow and Ice Masses. SC

Colbeck (Ed). Academic Press: New York, U.S.A.; pp 305 – 395.

Mann, G.W., Anderson, P.S. and Mobbs, S.D. (2000). Profile measurements of blowing snow at Halley, Antarctica. J. Geophys. Res., 105(D19), 24491–24508.

Page 108: Multiscale modelling of snow depth over an agricultural

90

Marsh, P. (1990) “Snow Hydrology”. In Northern Hydrology – Canadian Perspectives. Prowse

TD, Ommanney CSL (Eds). NHRI Science Report No. 1: Environment Canada, Saskatoon; pp

37 - 62.

Marshall, H.P., G. Koh, and R. Forster, (2004). CLPX-Ground: Ground-Based Frequency

Modulated Continuous Wave (FMCW) radar. National Snow and Ice Data Center Digital

Media, Boulder, CO, USA. http://nsidc.org/data/nsidc-0164.html.

McCuen, R. H., Leahy, R. B., and Johnson, P. A. (1990) Problems with logarithmic

transformations in regression. Journal Hydraulic Engineering, 116: 3, 414–428.

McMuen, R.H., Knight, Z., Cutter, A.G. (2006) Evaluation of the Nash-Sutcliffe Efficiency

Index, Journal of Hydraulic Engineering, 11: 6, 597-602, DOI: 10.1061/(ASCE)1084-

0699(2006)11:6(597)

McEwan, I. K. (1993). Bagnold‟s kink: a physical feature of a wind velocity profile modified by

blown sand?, Earth Surf. Process. Landf. 18, 145-156.

Nash, J.E., Sutcliffe, J.V. (1970) River flow forecasting through conceptual models. Part 1 – A

discussion of principles. Journal of Hydrology, 10: 282-290

Nelson, F.E., Hinkel, K.M., Shiklomanov, N.I., Mueller, G.R., Miller, L.L. and Walker, D.K.

(1998). Active-layer thickness in north central Alaska: Systematic sampling, scale, and spatial

autocorrelation. J. Geophys. Res., 103 (D22), 28963–28973.

NWS, cited (1996) Snow measurement guidelines. NOAA/NWS, Silver Springs, MD.

http://www.noaa.gov/os/coop/snowguid.htm

Oke, T.R. (1987) “Boundary Layer Climates”2nd

ed. Routledge,

Oke, T.R. and Hannell, F.G., (1966) “Variations of temperature within a soil” Weather 21. pp

21-8

Pietroniro, A et al. (2006) “Using the MESH modelling system for hydrological ensemble forecasting of the Laurentian Great Lakes at the regional scale” Hydrology and Earth System Sciences, No. 0000: 0001-15

Pomeroy, J.W., (1988) Wind Transport of Snow, Ph. D. Dissertation, University of Saskatchewan, Saskatoon, pp 226 Pomeroy, J.W., (1989) “A process-based model of snow drifting” Ann Glaciol., 13, pp. 237-240 Pomeroy, J.W. and Gray, D. M. (1990). Saltation of snow, Warter. Res. Res., 26 (7), 1583-1594. Pomeroy, J. W., Gray, D.M. and Landine, P.G. (1993). The Prairie Blowing Snow Model: characteristics, validation, operation. J. Hydrol., 144, 165–192.

Page 109: Multiscale modelling of snow depth over an agricultural

91

Pomeroy, J.W. et al, (2007b) “The Cold Regions Hydrological Model, a Platform for Basing

Process Representation and Model Structure on Physical Evidence” Hydrological Processes, 21.

No.19: 2650- 2667.

Pomeroy, J.W. and Gray, D. M. (1994). Sensitivity of snow relocation and sublimation to

climate and surface vegetation in Jones, H. G.. Davies, T. D., Ohmura, A,, and Morris, E. M.

(Eds), Snow and Ice Covers: Interactions with the Atmosphere and Ecosystems.

IAHS Publ. No. 223. Wallingford. pp. 213-225. Pomeroy, J.W. and Gray, D.M. (1995). “Snowcover Accumulation, Relocation and Management, National Hydrology Research Institute Science Report No. 7”, Hydrological Sciences Division, NHRI Division of Hydrology, University of Saskatchewan, Saskatoon. Pomeroy, J.W. and Male, D. H. (1986). Physical modelling of blowing snow for agricultural

production, in Steppuhn, H. And Nicholaichuk, W. (Eds), Proceedings Symposium of Snow

Management for the Great Central Plains Agriculture Council. Publ. No. 120, Water Studies

Institute, Saskatoon, Saskatchewan. pp. 73- 108

Pomeroy, J.W. and Male, D. H. (1992). Steady-state suspension of snow. J. Hydrol., 136, 275--

301. Pomeroy, J.W., Marsh, P. and Gray, D.M. (1997). Application of a distributed blowing snow model to the Arctic. Hydrol. Process., 11, 1451–1464. Pullum, G.K. (1991) The great Eskimo vocab hoax and other irreverent essays on the study of

language, The University of Chicago Press, pp. 160

Purves, R.S.. J.S. Barton, W.A. Mackaness and D.E. Sugden. (1998) The development of a rule-based spatial model of wind transport and deposition of snow. Ann. Glaciol., 26, 196-202 Rechard, P.A. and Larson, L.W. (1971). “The use of snow fences for shielding precipitation gauges”. Proc. 39th Western Snow Conference, Billings, Montana, pp. 56–62. Roeckner, E., et al., (1992) “Simulation of the present day climate with the ECHAM model:

Impact of model physics and resolution.” Max Planck Institute for Meteorology Rep. 93: 171

Roeckner, E. et al., (1996) “The atmospheric general circulation model ECHAM-4: Model

description and simulation of present day climate.” Max Planck Institute for Meteorology Rep.

218: 90.

Roeckner, E. et al., (2003) “The atmospheric general circulation model ECHAM5: Part I.” Max

Planck Institute for Meteorology Rep. 349: 127

Rutter, N. et al. (2009) “Evaluation of forest snow processes models (SnowMIP2)” J. Geophys.

Res., 114, D06111, doi:10.1029/2008JD011063

Page 110: Multiscale modelling of snow depth over an agricultural

92

Ryan, W. Et al. (2008) “Evaluation of Ultrasonic Snow Depth sensors for U.S. Snow

Measurements” Journal of Atmospheric and Oceanic Technology., pp. 667-684,

DOI: 10.1175/20077JTECHA947.1

Rykiel, E.J., Jr. (1996) Testing ecological models: the meaning of validation. Ecological

Modelling, 90, 229-244

Schmidt, R.A. (1972). “Sublimation of wind-transported snow - a model”. Research Paper RM-90, Rocky Mtn. Forestry and Range Expr. Sta., Forestry Service, US Department of Agriculture, Fort Collins, CO. Schmidt, R.A. (1982). Vertical profiles of wind speed, snow concentration, and humidity in blowing snow. Boundary-Layer Meteorol., 23, 223–246. Schmidt, R.A. (1986). Transport rate of drifting snow and the mean wind speed profile. Boundary-Layer Meteorol. 34, (3) 213-241 Schmidt, R.A. (1991). Sublimation of snow intercepted by an artificial conifer. Agric. Forest Meteorol., 54, 1–27. Schmidt, R.A. Jr, Gluns, D.R. (1991) “Snowfall interception on branches of three conifer

species”. Canadian Journal of Forest Research. 21: pp 1262 – 1269.

Shiklomanov, A.I., Lammers, R.B. and Vorosmarty, C.J. (2002). Widespread decline in hydrological monitoring threatens Pan-Arctic research. EOS Trans, Am. Geophys. Union, 83(2), 13, 16, 17. Spitalnic, S. et al. (1996) “An association between snowfall and ED presentation of cardiac arrest.” The American Journal of Emergency Medicine, vol 14, issue 6, pp 572-573

Singh, V.P., (1995) Computer Models of Watershed Hydrology. Water Resources Publication

Sturges, D. L., (1983) Establishment and growth of shelterbelt species and grass barriers on

windswept Wyoming rangeland. Transportation Research Record, 93 pp 4-11

Sturm M, Holmgren J, Liston GE. (1995) A seasonal snow cover classification system for local

to global applications. Journal of Climate, 8 1261-1283.

Sturm, M., Liston, G.E., Benson, C.S. and Holmgren, J. (2001a). Characteristics and growth of a snowdrift in arctic Alaska. Arctic, Antarctic Alpine Res., 33(3), 319–332. Sturm, M. and Liston, G.E. (2003). The snow cover on lakes of the Arctic Coastal Plain of Alaska, U.S.A. J. Glaciol., 49(166), 370–380.

Page 111: Multiscale modelling of snow depth over an agricultural

93

Sundsbø, P.-A. (1997) “Numerical modelling and simulation of snow drift” Ph.D. thesis, Norwegian University of Science and Technology, Trondheim; Narvik Institute of Technology, Department of Building Science, Narvik

Tabler, R.D. (1975). “Estimating the transport and evaporation of blowing snow.” In: Snow Management on Great Plains Symp., Bismarck, N. Dakota, July 1975, Proc. Great Plains Agric. Counc. Publ. vol 73, pp. 85–104. Tabler, R. D. (1980a) Geometry and density of drifts formed by snow fences. Journal of Glaciology, 26(94), pp 405-419 Takeuchi, M. (1980) Vertical profiles and horizontal increase of drift snow transport, J. Glaciol.,

26, 481-492.

Tarboton, D. G., T. G. Chowdhury and T. H. Jackson, (1995) “A Spatially Distributed Energy Balance Snowmelt Model”. In K. A. Tonnessen, M. W. Williams and M. Tranter (Ed.), Proceedings of a Boulder Symposium, July 3-14, IAHS Publ. No. 228: 141-155.

UNESCO/IHP (2009) “International Classification of Seasonal Snow on the Ground” IHP-VII

Technical Documents in Hydrology N° 83-IACS Contribution N° 1, UNESCO Working Series

SC-2009/WS/15

Verseghy, D.L. (2000) “Canadian Land Surface Scheme (CLASS): Its History and Future” Atmosphere-Ocean, 38. No. 1: 1-13 Walmsley, J. L., P. A. Taylor, and T. Keith, (1986) A simple model of neutrally stratified boundary-layer flow over complex terrain with surface roughness modulations (MS3DJH/3R). Boundary-Layer Meteorology, 36, pp. 157-224 Willmott, C.J., Matsuura, K, (2005) Advantages of the mean absolute error (MAE) over the root

mean square error (RMSE) in assessing average model performance. Climate Research 30:79-82 Yang, D., Goodison, B.E., Metcalfe, J.R., Golubev, V.S., Bates, R., Pangburn, T. and Hanson, C.L. (1998). Accuracy of NWS 800 standard nonrecording precipitation gauge: results and application of WMO intercomparison. J. Atmospher. Oceanic Tech., 15, 54–68. Yang, D., Kane, D.L., Hinzman, L.D., Goodison, B.E., Metcalfe, J.R., Louie, P.Y.T., Leavesley, G.H., Emerson, D.G. and Hanson, C.L. (2000). An evaluation of the Wyoming gauge system for snowfall measurement. Water Res. Res., 36, 2665–2677. Yang, D. and Ohata, T. (2001). A bias-corrected Siberian regional precipitation climatology. J. Hydrometeorol., 2, 122–139.

Page 112: Multiscale modelling of snow depth over an agricultural

94

Appendix 1

Parameter Definitions:

Basin Module

basin_area (km2) - basin area.

hru area (km2) - HRU area.

hru_lat (°) - latitude.

hru_elev (m) - altitude.

hru_GSL (°) - ground slope.

hru_ASL (°) - aspect.

basin_name – name given to the basin

hru_names – name given to the HRUs

run_ID - integer number.

Obs Module

basin_area (km2) - basin area.

hru area (km2) - HRU area.

hru_elev (m) - altitude.

obs_elev (m) - altitude.

lapse_rate (°C/100m) - lapse rate correction.

tmax_allrain (°C) - precipitation is all rain when the temperature is greater or equal to this

value.

tmax_allsnow (°C) - precipitation is all snow when the temperature is less or equal to this

value.

catchadjust () – 0 = for none 1 = Goodison

ppt_daily_distrib () – 0 = daily precip all in first interval, 1 = equally divided over the day

MSM_Org Module

a1 (s) - Albedo decay time constant for cold snow.

a2 (s) - Albedo decay time constant for melting snow.

amin () - Minimum albedo for aged snow.

amax () - Maximum albedo for fresh snow.

smin (mm) - Minimum snowfall to refresh snow albedo.

Z0snow (m) - snow roughness length.

Zref (m) - reference height.

basin_area (km2) - basin area.

hru area (km2) - HRU area.

hru_elev (m) - altitude.

PBSM Module

fetch (m) - fetch distance. (300-1000m)

Ht (m) - crop height.

Page 113: Multiscale modelling of snow depth over an agricultural

95

distrib () - distribution fractions. Value for HRU 1 controls snow transport into the basin.

N_S (1/m2) - vegetation number density.

A_S (m) - stalk diameter or silhouette.

basin_area (km2) - basin area.

hru_area (km2) - hru_area.

inhibit_evap (flag) - an output parameter set true when the SWE is greater than zero. It is

used to inhibit evaporation from the evaporation modules.

inhibit_bs (flag) - an input inhibiting blowing snow when set equal to 1. Inhibited HRU

is still able to receive drift from other HRUs.

Page 114: Multiscale modelling of snow depth over an agricultural

96

Appendix 2

Dec. 08/08

Jan. 15/09

Jan. 21/09

334

335

336

337

338

339

340

341

342

343

344

0

10

20

30

40

50

60

70

0.0

0

30

.46

62

.37

95

.89

12

8.3

6

15

7.5

1

18

2.7

7

21

3.4

2

23

9.6

9

26

5.4

7

29

1.8

1

31

9.3

4

34

9.1

2

37

9.7

0

41

3.5

8

43

8.4

9

46

6.4

2

49

1.1

8

52

0.1

5

55

3.5

1

58

3.4

7

61

5.1

8

64

3.3

4

67

2.7

0

69

9.7

3

72

3.6

2

74

9.6

3

77

4.7

2

Elev

atio

n (m

)

Sno

wd

epth

(cm

)

Distance (m)

Transect 1 - Dec. 08/08

334

335

336

337

338

339

340

341

342

343

344

345

0

5

10

15

20

25

30

35

40

45

0.00

26.2

3

56.8

6

87.9

6

119.

92

151.

37

171.

91

197.

34

229.

12

263.

27

296.

40

329.

10

361.

41

392.

64

422.

87

452.

97

484.

81

516.

80

547.

50

577.

82

610.

33

640.

77

673.

82

707.

76

739.

42

Ele

vati

on

(m)

Sno

wd

ep

th (c

m)

Distance (m)

Transect 1 - Jan. 15/09

334

336

338

340

342

344

346

0

10

20

30

40

50

60

70

80

90

0.00

20.9

649

.06

77.5

110

5.27

132.

4915

9.17

173.

0419

7.68

226.

4225

5.46

285.

3731

3.74

342.

3737

1.27

401.

1443

2.37

463.

5649

1.80

520.

8055

1.65

584.

4861

3.58

641.

2967

0.47

700.

0372

9.18

757.

09

Elev

atio

n (m

)

Sno

wd

ep

th (c

m)

Distance (m)

Transect 1 - Jan. 21/09

335

336

337

338

339

340

341

342

343

344

0

10

20

30

40

50

60

70

80

0.00

25

.09

50

.57

76

.86

101.

90

127.

77

153.

37

178.

50

204.

59

230.

80

249.

81

267.

02

290.

21

312.

52

335.

24

367.

92

392.

52

417.

56

441.

64

465.

51

489.

45

513.

08

537.

57

562.

08

586.

88

Elev

atio

n (m

)

Sno

wd

epth

(cm

)

Distance (m)

Transect 3 - Jan. 21/09

330

332

334

336

338

340

342

344

346

348

0

10

20

30

40

50

60

70

80

90

100

0.0

0

38

.14

73

.46

11

2.2

7

15

3.0

3

18

8.5

1

22

7.0

0

26

4.3

3

30

3.1

0

34

1.7

6

37

8.1

1

40

5.9

4

44

6.1

6

48

6.3

5

52

8.0

5

57

0.7

6

61

3.4

8

65

5.6

2

69

8.0

0

73

7.5

4

78

1.0

4

82

1.6

7

86

1.4

5

90

4.9

9

94

5.9

0

Elev

atio

n (m

)

Sno

wd

epth

(cm

)

Distance (m)

Transect 2 - Dec. 08/08

330

332

334

336

338

340

342

344

346

348

0

10

20

30

40

50

60

70

80

90

100

0.00

36.8

9

73.0

1

10

9.7

2

14

5.7

0

18

1.3

3

21

6.6

1

25

5.4

4

30

1.3

5

34

4.8

8

37

8.8

1

39

7.9

5

43

3.1

6

46

8.2

6

50

5.8

7

54

4.5

1

58

3.9

9

62

4.2

2

66

2.2

2

70

0.7

2

73

7.6

5

77

6.9

0

83

1.7

9

87

0.9

3

91

3.3

5

95

6.6

8

Ele

vati

on

(m)

Sno

wd

ep

th (c

m)

Distance (m)

Transect 2 - Jan 15/09

332

334

336

338

340

342

344

346

348

0

10

20

30

40

50

60

70

0.00

38.2

6

75.5

3

118.

07

160.

13

202.

78

243.

27

284.

18

324.

90

365.

76

389.

44

424.

91

461.

90

498.

19

534.

85

572.

98

610.

13

645.

80

683.

66

721.

17

758.

99

810.

67

850.

36

886.

61

921.

61

958.

20

Elev

atio

n (m

)

Sno

wd

ep

th (c

m)

Distance (m)

Transect 2 - Jan. 21/09

332

334

336

338

340

342

344

346

0

20

40

60

80

100

120

0.00

23.3

9

48.7

3

71.7

7

95.3

4

119.

70

132.

99

143.

10

165.

57

187.

83

211.

15

236.

33

260.

89

285.

12

310.

06

334.

29

357.

33

381.

04

406.

25

430.

67

454.

81

Elev

atio

n (m

)

Sno

wd

epth

(cm

)

Distance (m)

Transect 4 - Jan. 21/09

Page 115: Multiscale modelling of snow depth over an agricultural

97

Jan. 29/09

340

341

342

343

344

345

346

347

348

0

10

20

30

40

50

60

70

80

90

100

0.0

0

8.3

7

17

.07

26

.54

35

.81

47

.69

56

.74

66

.85

76

.22

85

.29

95

.53

10

4.3

9

11

4.5

2

12

4.1

7

13

5.2

3

14

7.1

8

15

7.8

4

16

9.7

9

17

9.9

6

19

0.4

6

20

1.4

0

21

2.2

9

22

3.6

5

23

4.9

5

24

6.7

3

25

9.1

9

27

0.4

6

Elev

atio

n (m

)

Sno

wd

epth

(cm

)

Distance (m)

Western Transect - Jan. 21/09

338

339

340

341

342

343

344

345

346

0

5

10

15

20

25

30

35

40

45

0.00

25.9

9

52.1

5

77.4

7

103.

20

128.

65

154.

16

176.

82

200.

44

224.

62

249.

76

275.

15

300.

42

326.

03

352.

70

380.

45

405.

43

431.

54

458.

65

484.

84

510.

67

536.

34

561.

87

590.

90

614.

46

Ele

vati

on

(m)

Sno

wd

ep

th (c

m)

Distance (m)

Northern Transect - Jan. 21/09

330

332

334

336

338

340

342

344

346

348

0

10

20

30

40

50

60

70

80

0.00

41.2

4

85.2

6

127.

77

170.

18

213.

39

249.

81

282.

78

320.

04

367.

92

408.

27

449.

71

489.

45

529.

37

570.

44

611.

38

652.

60

695.

16

737.

09

778.

17

818.

73

858.

98

899.

02

Ele

vati

on

n (m

)

Sno

wd

ep

th (c

m)

Distance (m)

Southern Transect - Jan. 21/09

336

337

338

339

340

341

342

343

344

345

346

347

0

20

40

60

80

100

120

140

0.00

34.1

0

65.5

2

98.1

8

131.

58

154.

54

173.

16

206.

35

239.

41

273.

86

307.

51

341.

72

375.

95

409.

67

443.

57

478.

76

513.

97

547.

34

584.

16

620.

03

654.

84

695.

91

728.

95

759.

91

Elev

atio

n (m

)

Sno

wd

epth

(cm

)

Distance (m)

Transect 1 - Jan. 29/09

337

338

339

340

341

342

343

344

345

0

10

20

30

40

50

60

70

80

90

0.0

0

17

.11

34

.24

51

.58

68

.69

86

.76

10

3.7

8

12

1.3

6

13

9.4

2

15

6.5

7

17

5.1

6

19

1.5

1

20

9.3

0

22

6.7

9

24

4.0

7

26

1.5

8

27

9.1

0

29

6.2

2

30

3.9

3

31

2.4

0

32

8.9

1

34

5.6

5

Elev

atio

n (m

)

Sno

wd

epth

(cm

)

Distance (m)

Northwestern (Road) Transect - Jan. 21/09

342

342.5

343

343.5

344

344.5

345

345.5

346

346.5

0

10

20

30

40

50

60

70

80

90

100

0.00

16.9

6

32.9

9

49.5

4

66.7

8

83.8

0

101.

04

116.

98

134.

61

151.

40

168.

67

186.

63

204.

82

222.

32

239.

36

256.

35

273.

71

291.

21

308.

84

325.

96

342.

55

359.

61

376.

10

391.

47

407.

97

425.

52

442.

71

Ele

vati

on

(m)

Sno

wd

ep

th (c

m)

Distance (m)

Eastern Transect - Jan. 21/09

332

334

336

338

340

342

344

346

348

0

20

40

60

80

100

120

0.00

38.0

4

77.4

3

117.

75

161.

00

209.

05

252.

08

294.

94

338.

63

377.

85

400.

48

437.

29

474.

80

513.

11

555.

16

593.

42

633.

44

673.

16

716.

15

762.

80

804.

99

848.

16

885.

98

923.

00

960.

86

Elev

atio

n (m

)

Sno

wd

epth

(cm

)

Distance (m)

Transect 2 - Jan. 29/09

Page 116: Multiscale modelling of snow depth over an agricultural

98

332

334

336

338

340

342

344

346

0

10

20

30

40

50

60

70

80

90

0.0

0

26

.38

56

.23

81

.77

10

6.1

9

13

0.5

6

14

3.3

6

16

9.2

6

19

3.3

6

21

9.9

2

24

4.1

3

26

7.8

3

29

4.3

3

31

9.2

3

34

6.2

0

36

9.3

0

39

3.9

0

41

8.4

0

44

3.1

7

Elev

atio

n (m

)

Sno

wd

epth

(cm

)

Distance (m)

Transect 3 - Jan. 29/09

342

343

344

345

346

347

348

349

0

20

40

60

80

100

120

0.00

16

.84

29

.56

43

.31

58

.16

74

.67

91

.12

108.

04

125.

98

143.

08

161.

56

178.

90

196.

11

213.

29

225.

60

240.

20

256.

30

271.

45

287.

31

307.

26

Elev

atio

n (m

)

Sno

wd

epth

(cm

)

Distance (m)

Western Transect - Jan. 29/09

338

339

340

341

342

343

344

345

0

5

10

15

20

25

30

35

40

45

0.00

25.5

1

52.5

3

80.5

2

107.

61

133.

44

158.

84

183.

38

208.

97

234.

80

261.

10

288.

06

314.

55

340.

16

366.

58

392.

93

418.

15

445.

26

471.

94

496.

92

524.

34

550.

60

579.

81

605.

48

Ele

vati

on

(m)

Sno

wd

ep

th (c

m)

Distance (m)

Northern Transect - Jan. 29/09

330

332

334

336

338

340

342

344

346

348

0

10

20

30

40

50

60

0.00

46.0

7

90.9

3

134.

68

177.

20

221.

45

258.

15

295.

66

335.

54

377.

69

419.

10

457.

02

496.

84

536.

60

576.

45

618.

04

660.

07

702.

96

746.

43

788.

22

827.

84

866.

87

907.

32

Elev

atio

n (m

)

Sno

wd

epth

(cm

)

Distance (m)

Southern Transect - Jan. 29/09

336

337

338

339

340

341

342

343

344

345

0

20

40

60

80

100

120

0.0

0

26

.34

51

.87

77

.84

10

2.7

8

12

6.8

5

14

4.3

3

16

8.8

4

19

1.1

7

21

3.1

9

23

8.1

6

25

8.8

8

28

0.0

1

29

9.9

2

32

2.2

0

34

5.1

9

36

6.3

8

38

5.4

6

40

7.3

4

42

9.4

4

45

1.3

8

47

3.7

4

49

8.7

3

52

2.5

0

54

4.7

6

Elev

atio

n (m

)

Sno

wd

epth

(cm

)

Distance (m)

Transect 4 - Jan. 29/09

336

337

338

339

340

341

342

343

344

345

0

10

20

30

40

50

60

70

80

0.00

16.7

4

33.0

3

50.6

8

66.9

2

83.7

4

99.5

6

115.

10

132.

67

149.

07

166.

25

184.

83

200.

88

217.

43

233.

95

249.

44

264.

46

281.

39

298.

05

308.

11

322.

63

338.

14

354.

17

Ele

vati

on

(m)

Sno

wd

epth

(cm

)

Distance (m)

Northwestern (Road) Transect - Jan. 29/09

341

342

343

344

345

346

347

0

10

20

30

40

50

60

70

80

0.00

23.9

8

46.6

1

70.2

6

92.7

4

115.

60

136.

81

158.

76

181.

96

205.

34

229.

36

252.

06

274.

83

298.

97

324.

78

350.

34

372.

90

396.

14

421.

76

Ele

vati

on

(m)

Sno

wd

ep

th (c

m)

Distance (m)

Eastern Transect - Jan. 29/09

Page 117: Multiscale modelling of snow depth over an agricultural

99

Feb. 04/09

335

336

337

338

339

340

341

342

343

344

345

0

20

40

60

80

100

120

0.0

0

33

.36

67

.17

10

0.7

3

13

3.4

9

15

1.1

5

18

5.4

0

22

1.1

1

25

6.3

3

29

1.1

1

32

6.0

8

36

0.0

4

39

4.9

8

43

0.9

4

46

7.2

5

50

0.5

8

53

3.9

2

57

1.7

0

60

6.7

9

64

1.8

5

67

6.5

4

71

2.2

4

74

4.9

2

Elev

atio

n (m

)

Sno

wd

epth

(cm

)

Distance (m)

Transect 1 - Feb. 04/09

332

334

336

338

340

342

344

0

10

20

30

40

50

60

70

80

90

100

0.00

16.7

2

33.6

3

54.1

7

71.2

4

87.9

4

103.

83

119.

18

129.

19

141.

39

157.

49

172.

79

190.

78

208.

54

226.

04

243.

89

261.

91

281.

06

299.

65

317.

83

335.

01

352.

33

370.

07

387.

17

404.

79

422.

90

439.

97

Ele

vati

on

(m)

Sno

wd

ep

th (c

m)

Distance (m)

Transect 3 - Feb. 04/09

340

341

342

343

344

345

346

347

0

10

20

30

40

50

60

70

80

90

100

Ele

vati

on

(m)

Sno

wd

ep

th (c

m)

Distance (m)

Western Transect - Feb. 04/09

337

338

339

340

341

342

343

344

0

5

10

15

20

25

30

35

40

45

50

0.00

28

.15

54

.47

81

.34

108.

71

136.

47

162.

89

189.

82

216.

63

244.

14

270.

19

296.

07

320.

51

347.

56

373.

79

399.

20

426.

26

453.

91

480.

28

507.

90

535.

12

562.

55

589.

11

615.

60

Elev

atio

n (m

)

Sno

wd

epth

(cm

)

Distance (m)

Northern Transect - Feb. 04/09

330

332

334

336

338

340

342

344

346

0

20

40

60

80

100

120

140

0.0

0

38

.51

78

.11

12

0.5

9

16

7.0

9

21

8.1

2

26

7.5

6

31

1.5

4

35

7.6

6

39

0.0

2

42

3.2

1

46

2.6

7

50

3.9

2

54

6.1

8

58

9.0

9

62

8.7

7

67

0.1

3

71

4.9

9

75

9.2

9

80

5.8

5

86

5.3

9

90

7.7

7

94

8.9

9

Elev

atio

n (m

)

Sno

wd

epth

(cm

)

Distance (m)

Transect 2 - Feb. 04/09

335

336

337

338

339

340

341

342

343

344

0

10

20

30

40

50

60

70

80

90

0.00

25.6

6

51.6

6

78.5

2

105.…

132.…

147.…

171.…

198.…

223.…

248.…

272.…

298.…

324.…

351.…

374.…

404.…

429.…

454.…

480.…

507.…

532.…

557.…

Ele

vati

on

(m)

Sno

wd

ep

th (c

m)

Distance (m)

Transect 4 - Feb. 04/09

334

335

336

337

338

339

340

341

342

343

344

345

0

10

20

30

40

50

60

70

80

90

100

0.00

16.6

8

32.7

7

48.7

0

66.6

1

83.8

3

100.

63

117.

26

135.

05

152.

01

172.

46

188.

36

205.

75

221.

71

239.

58

256.

78

271.

86

288.

15

302.

07

312.

98

330.

52

346.

68

Ele

vati

on

(m)

Sno

wd

ep

th (c

m)

Distance (m)

Northwestern (Road) Transect - Feb. 04/09

339.5

340

340.5

341

341.5

342

342.5

343

343.5

344

344.5

0

10

20

30

40

50

60

70

80

0.00

21.7

2

41.4

9

65.0

5

94.1

1

119.

94

145.

14

170.

48

195.

77

220.

25

244.

00

266.

68

287.

96

309.

01

331.

58

356.

26

382.

78

408.

42

434.

80

Elev

atio

n (m

)

Sno

wd

epth

(cm

)

Distance (m)

Eastern Transect - Feb. 04/09

Page 118: Multiscale modelling of snow depth over an agricultural

100

328

330

332

334

336

338

340

342

344

346

0

10

20

30

40

50

60

70

80

90

0.0

0

35

.63

71

.01

10

5.5

5

13

8.9

8

17

3.6

9

21

4.8

3

25

0.5

2

28

3.7

6

31

3.0

8

34

5.3

3

38

0.1

6

41

3.0

1

44

7.6

9

48

2.8

0

51

9.2

8

55

3.3

9

59

0.0

4

62

6.7

4

66

2.7

4

69

9.8

2

73

5.0

3

77

3.7

5

81

3.9

8

85

5.3

4

89

1.1

1

92

4.8

7

Elev

atio

n (m

)

Sno

wd

epth

(cm

)

Distance (m)

Southern Transect - Feb. 04/09